Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Jul 22.
Published in final edited form as: J Am Chem Soc. 2022 Apr 26;144(17):7517–7530. doi: 10.1021/jacs.1c07990

Beyond Triphosphates: Reagents and Methods for Chemical Oligophosphorylation

Scott M Shepard , Henning J Jessen , Christopher C Cummins
PMCID: PMC9307064  NIHMSID: NIHMS1822270  PMID: 35471019

Abstract

Oligophosphates play essential roles in biochemistry, and considerable research has been directed towards the synthesis of both naturally occurring oligophosphates and their synthetic analogs. Greater attention has been given to mono-, di-, and triphosphates, as these are present in higher concentrations biologically and easier to synthesize. However, extended oligophosphates have potent biochemical roles, ranging from blood coagulation to HIV drug resistance. Sporadic reports have slowly built a niche body of literature related to the synthesis and study of extended oligophosphates, but newfound interests and developments have the potential to rapidly expand this field. Here we report on current methods to synthesize oligophosphates longer than triphosphates and comment on the most important future directions for this area of research. The state of the art has provided fairly robust methods for synthesizing nucleoside 5’-tetra- and pentaphosphates as well as dinucleoside 5’,5’-oligophosphates. Future research should endeavor to push such syntheses to longer oligophosphates while developing synthetic methodologies for rarer morphologies such as 3’-nucleoside oligophosphates, polyphosphates, and phosphonate/thiophosphate analogs of these species.

Graphical Abstract

graphic file with name nihms-1822270-f0001.jpg

Introduction

Chemical synthesis and the study of biological systems have naturally progressed in tandem, with synthesized biomolecules and their non-natural analogs playing a key role in many biochemical experiments.1 Accordingly, whole fields of synthetic chemistry have arisen around the synthesis of preeminent classes of biomolecules such as peptides,24 oligonucleotides,57 carbohydrates,8,9 etc. Phosphorus, as an essential element for all life,10 is necessarily included in these synthetic pursuits, especially in the study of nucleotides, DNA/RNA, and protein phosphorylation. However, in contrast to the rich field of carbon bond centric organic chemistry, biorelevant phosphorus chemistry is less well developed. Phosphorus exists in biological systems almost exclusively as P(V) phosphates,11 and phosphate synthesis is mostly limited to simple nucleophilic substitution reactions, despite the frequent departure from biological conditions to utilize P(III) chemistry. Typically, phosphorus bound leaving groups are substituted with a nucleophilic substrate such as an amine, alcohol, or phosphate of interest (Figure 1, Top). This general methodology is simple but powerful enough to achieve transformations as difficult as oligonucleotide synthesis.5 However, one of the most important properties of phosphates is their ability to form longer linear or cyclic oligophosphates,12 such as adenosine triphosphate (ATP) and other nucleoside 5’-triphosphates (NTPs). The synthesis of oligophosphates has remained an enduring challenge, and selectively synthesizing extended oligophosphates requires pushing the limits of phosphorylation chemistry.

Figure 1.

Figure 1.

Top: A generalized phosphorylation reaction in which a nucleophile substitutes a leaving group from an electrophilic phosphate. Bottom: Examples of extended oligophosphates present in biological systems: nucleoside 5’-oligophosphates (pnN), dinucleoside 5’,5’-oligophosphates (NpnN), and inorganic polyphosphate (PolyP).

ATP is perhaps the most important and well known biological cofactor, acting as the primary mediator of energy transfer in cellular processes.13 ATP and the other NTPs are also the monomers for biological DNA/RNA synthesis14 as well as signaling molecules,15 resulting in paramount utility of these compounds and their analogs for biochemical studies and drug candidates.16,17 Accordingly, the chemical synthesis of NTPs has received considerable attention, beginning with the first synthesis of ATP by Khorana in 1954 through N,N’-dicyclohexylcarbodiimide (DCC) mediated coupling of adenosine monophosphate (AMP) and orthophosphate.18 Since then, considerable research effort has been devoted to developing improved syntheses of NTPs, resulting in an at times confusing variety of reaction conditions and methods.17,19 However, NTPs are far from the only oligophosphates utilized by Nature. Nucleoside 5’ tetra- and pentaphosphates (Np4 and Np5)2022 have been identified in biological sources as well as more complex morphologies such as dinucleoside 5’,5’-oligophosphates (NpnNs)23,24 (Figure 1, Bottom). Some of these compounds and their analogs have been known for decades and even investigated as therapeutics in clinical trials,25 but the difficulty of isolating such low abundance compounds from natural sources and dearth of synthetic methodology has hampered their study. Phosphates are also found in biology in the form of long inorganic polyphosphates (PolyP, up to 100s of phosphate units long)26 and more recently, as oligo-27 and polyphosphorylated peptides28 (Figure 1, Bottom). While many lessons and methodologies from NTP synthesis can be applied to extended oligophosphates, research on the synthesis of these molecules is much rarer.

In this Perspective, we will review the established and cutting edge methods for chemical oligophosphorylation with an emphasis on the synthesis of oligophosphates longer than triphosphates. As an oligophosphorylation reagent or method can be utilized to generate products with a variety of different phosphate chain lengths, this Perspective is structured into sections highlighting methodologies for appending different numbers of phosphate units, that is monophosphorylation, diphosphorylation, etc. Although not the prime focus of this Perspective, the applications, biological and otherwise, of extended oligophosphates of defined length, as well as PolyP that contains a distribution of chain lengths, will be discussed briefly.

General Considerations and Purification

Phosphates, as oxyanions, are encountered either as salts, neutral acids, or a combination thereof with major consequences for their reactivity. In aqueous systems, these salts are commonly those of alkali metal or ammonium (NH4) cations. However, many of the activating agents used in synthetic phosphorylation chemistry are either water reactive or insoluble in water. Furthermore, electrophilic phosphorus species are often unstable with respect to water as well. Therefore, phosphorylation reactions conducted in water are usually limited to enzymatic reactions,29,30 certain “green” reactions,31 some reactions of phosphorimidazolides,32,33 and those using specialized water soluble reagents such as 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDAC).34 Synthetic phosphorylation chemistry is typically conducted in polar organic solvents such as acetonitrile, dimethylformamide (DMF), and pyridine, requiring the use of lipophilic cations. For these transformations, phosphates are often encountered as tetrabutylammonium (TBA) or less soluble tributylammonium salts. These lipophilic salts are readily accessed by first acidifying the desired phosphate compound with an acidic resin, such as Dowex 50WX8, and then neutralization with either TBA hydroxide or tributylamine, followed by lyophilization.35 The bis(triphenylphosphine)iminium (PPN) cation has also been used and typically gives more crystalline phosphate salts.36 The nucleophilicity of phosphates is enhanced when they are deprotonated, and nucleotide substrates are commonly encountered fully deprotonated, such as the salt [TBA]2[AMP].35 However, especially in the absence of hydrogen bond donor groups, retention of some protons is necessary for stability in organic solvents. For example, [TBA]3[HP2O7] is readily prepared and soluble in polar organic solvents, but [TBA]4[P2O7] is accessible only in water.

Most synthetic phosphorylation chemistry can be divided into reactions that utilize P(III) reagents and those that operate in an entirely P(V) regime. Prime motivations for P(III) chemistry are the greater reactivity of these reagents and facile access to thiophosphate products. However, in addition to requiring air free conditions, the great reactivity of P(III) reagents often, although not always,37,38 necessitates the use of protecting groups, for example of nucleoside 2’ and 3’ positions. In a typical example, an electrophilic P(III) compound bearing a leaving group is synthesized, ultimately from phosphorus trichloride. Treatment with a nucleophile, followed by oxidation or sulfurization yields the desired phosphate or thiophosphate product (Figure 2, top). P(V) methods usually start with a phosphate of interest that is then activated to form an electrophilic compound. Common activating agents are sulfonyl chlorides,35 carbodiimides,18 carboxylic acid anhydrides,39 and carbonyldiimidazole (CDI).40 Importantly, phosphates contain multiple nucleophilic oxygen positions, and treating an inorganic phosphate with an activating agent results in oligomerization, especially cyclic metaphosphates.41,42 Similarly, treatment of a substituted triphosphate with an activating agent results in an intramolecular dehydration to form a substituted trimetaphosphate (TriMP) compound (Figure 2, bottom). However, the resulting metaphosphate is still a powerful electrophile due to the phosphotriester functionality, and ring opening of these intermediates has been used repeatedly in the synthesis of oligophosphates.35,37,38,4346

Figure 2.

Figure 2.

Top: General P(III) phosphorylation conditions in which a P(III) electrophile is treated with a nucleophile followed by oxidation or sulfurization. Bottom: General methods for electrophilic activation of P(V) phosphates in which a mono- or diphosphate is treated with an activating agent to generate a leaving group, or an extended oligophosphate is treated with an activating agent to generate an electrophilic substituted metaphosphate.

The purification of oligophosphates is often the most difficult and expensive part of a synthetic procedure. As mixtures of phosphate salts with similar solubilities, purification of crude oligophosphates by crystallization alone is extremely rare. The standard purification method is chromatography, typically High Performance Liquid Chromatography (HPLC) for small scales and flash or medium pressure Liquid Chromatography (LC) for larger scales. Traditional reverse phase (RP) chromatography is a poor separation method for most biological oligophosphates, as these highly polar compounds elute rapidly. In contrast, anion exchange (AX) chromatography strongly retains anionic oligophosphates and is the most common purification method.47 Ion pairing RP-HPLC48 and hydrophilic interaction chromatography (HILIC)49 methods are also sometimes used for purification. Electrophoresis methods, such as Capillary Electrophoresis (CE)50,51 and Polyacrylamide Gel Electrophoresis (PAGE),52 can also show good efficiency for oligophosphate separation, but they are not typically applied on preparative scales. A key target in continuing phosphorylation chemistry is the development of high yielding reactions and other methods to avoid or simplify costly chromatographic separations.

Mono-, Di-, and Triphosphates

The many diverse biological roles of monophosphates, diphosphates, and triphosphates is a vast topic that is intimately intertwined with all aspects of biochemistry. Extremely briefly, such phosphates mediate energy transfer, enzyme regulation, DNA/RNA synthesis, signal transduction, bone growth, and much more far beyond the scope of this Perspective.53 Nevertheless, it behooves us to examine the general synthetic methodologies used to access these species and how these relate to the synthesis of extended oligophosphates. The synthesis of nucleoside mono-, di-, and triphosphates has also been reviewed several times.17,19,54

Synthesis of monophosphates is limited to coupling a desired organic substrate (such as a nucleoside) and a monophosphate synthon (either P(III) or P(V)). In a simple example, the Yoshikawa method, nucleosides are treated with phosphoryl chloride in alkylphosphate solvents to selectively phosphorylate the 5’ position (Figure 3).55 In a much more complex example, P(III) chemistry is used in the phosphoramidite method for the synthesis of oligonucleotides (Figure 3).7 In this method, the 5’-hydroxyl position of a silica supported oligonucleotide is coupled with a P(III) 3’-phosphoramidite nucleoside bearing a 4,4’-dimethoxytrityl (DMT) protecting group on the 5’ position and a 2-cyanoethyl protecting group on the phosphoramidite oxygen. After coupling, the product is capped with acetic anhydride and a nucleophilic base to prevent chain elongation of any remaining unreacted 5’ hydroxyl groups. The phosphorus center is then oxidized, and the DMT group is cleaved with acid, providing a free 5’ hydroxyl group for further oligonucleotide chain extension. After the full oligonucleotide is synthesized, basic conditions cleave the molecule from the silica support and deprotect the 2-cyanoethyl groups. A related method of oligonucleotide synthesis that allows incorporation of thiophosphate moieties has also been introduced recently by Baran.56,57

Figure 3.

Figure 3.

Representative syntheses of mono-7,55 and diphosphates.40,58

In principle, the synthesis of diphosphates can be performed either by coupling an organic substrate and a diphosphate or by monophosphorylation of a monophosphate substrate. The former is analogous to monophosphorylation, but diphosphate electrophiles are rare. An example of such a species is a mixed P(III)-P(V) methylene bridged analog of pyrophosphate used by Filipov63 and Fiedler64 to synthesize bisphosphonates. In an unusual phosphorylation example, Poulter used pyrophosphate as a nucleophile to react with an electrophilic organic substrate, the opposite philicities of most phosphorylation chemistry (Figure 3).58 In contrast to scarce diphosphorylation chemistry, utilizing a monophosphorylation procedure with a monophosphate substrate is typically quite facile, due to the great nucleophilicity of monophosphates. For example, NMPs react with CDI to form electrophilic phosphorimidazolides. Subsequent reaction with another monophosphate substrate results in a diphosphate (Pankiewicz, Figure 3).40 Phosphorimidazolides have been used extensively for such chemistry,17 with notable improvements such as aqueous conditions and faster reactions due to added zinc salts.32

Similarly, the synthesis of triphosphates can be performed either by (1) coupling an organic substrate with a triphosphate source, (2) diphosphorylation of a monophosphate, or (3) monophosphorylation of a diphosphate. The first method is the most recent development, pioneered by Taylor,35,59,61,65,66 in which a nucleophilic substrate is treated with a leaving group modified trimetaphosphate (Figure 4). Notably, Taylor’s initial methodology relies on in situ activation of TriMP, but subsequent reports have described isolation and structural characterization of analogous activated TriMP species as crystalline compounds (Figure 4).61,62 Jessen37,38 and Huang67 have elaborated similar triphosphorylation approaches that instead utilize mixed P(III)-P(V) analogs of TriMP. The second route to triphosphates is the most classic, including Khorana’s original ATP synthesis,18 and the well known Ludwig-Eckstein synthesis of NTPs (Figure 4).60 These methods have significant drawbacks in terms of yield and purification, and Chaput has addressed these issues by utilizing a pyrene modified pyrophosphate reagent.68 In addition to a more efficient coupling step, the pyrene modification allows purification by normal phase chromatography, significantly improving the cost and scalability of purification. Lastly, the third approach is analogous to the monophosphorylation procedures previously discussed.69,70 In this manner, iterative monophosphorylation reactions can be used to convert monophosphates into diphosphates and then triphosphates, such as the work by Jessen71,72 and Chaput’s synthesis of TNA triphosphates.73 However, diphosphate substrates are much less nucleophilic than monophosphate substrates, often resulting in longer reaction times and lower yields.37

Figure 4.

Figure 4.

Representative syntheses of triphosphates.18,59,60 Thermal ellipsoid plots of crystallographically characterized triphosphorylation reagents61,62 ([PPN][P3O9P(NC4H8)3]·2H2O, [PPN][HNEt3][P3O9Ts]) with thermal ellipsoids set at 50% probability and hydrogen atoms, cations, and solvent molecules omitted for clarity (red for oxygen, orange for phosphorus, blue for nitrogen, grey for carbon, yellow for sulfur). The thermal ellipsoid plot of [P3O9Ts]2− was adapted from reference 62 as permitted under its CC BY 4.0 License https://creativecommons.org/licenses/by/4.0/.

Biological Synthesis of Oligophosphates

Although there are notable examples of enzymatic syntheses of oligophosphates,17,74,75 the nonaqueous synthetic chemistry focused on in this Perspective is a large departure from native biological syntheses of oligophosphates. However, the biological synthesis and utilization of these compounds is both a source of inspiration and a vital biochemical application for synthetic oligophosphates. Phosphorylation and dephosphorylation enzymes are broadly divided into phosphotransferases, phosphorylases, and phosphatases, representing an extremely broad range of enzymes.76 Kinases are perhaps the most familiar phosphotransferase subclass, catalyzing phosphoryl transfer from ATP to form phosphorylated substrates and ADP (Figure 5). This is analogous to synthetic monophosphorylation reactions, with ADP acting as a leaving group (Figure 5). The vast number (~500 known human kinases)77 and high selectivity of kinases indicate the need for careful preassociation of substrates and hydrogen bonding networks that enzyme active sites provide. Overall, these aqueous phosphorylation conditions are quite mild compared to the water reactive and highly electrophilic phosphorus reagents that are common in synthetic phosphorylation chemistry. Additionally, ATP, the most common source of biological phosphoryl equivalents, is almost entirely bound to magnesium under physiological conditions. This divalent cation is important for screening charge and facilitating phosphorylation reactions,78 a strategy that has been widely applied to difficult synthetic phosphorylations.32,35,37,46,79

Figure 5.

Figure 5.

Generalized reaction scheme of a kinase.

Although nucleotides are a stereotypical class of oligophosphates, biology also extensively utilizes inorganic polyP.26,80 These compounds have been known for decades across all kingdoms of life, particularly due to the presence of polyP granules in bacteria that are visible through microscopy. In bacteria, polyP is synthesized and utilized by polyphosphate kinases (PPK1 and PPK2) which transfer the terminal phosphate from ATP to polyP as well as the reverse reaction.81 A recent solid state NMR study backs up previous tentative reports of cyclic metaphosphates in these polyphosphate granules,82 but the potential biological importance of metaphosphates is an open question. PolyP is present in mammals as well and is involved in blood clotting.83,84 The biosynthesis of polyP in humans has been an enduring mystery.85 A recent report detailed synthesis of PolyP by F0F1-ATP synthase in mammalian mitochondria,86 and this result awaits further validation as the source of human polyP.

A few groups have developed preparative enzymatic syntheses of extended oligophosphates. As early as 1989, Tomita utilized Leucyl t-RNA synthetase for the synthesis of adenosine 5’-tetraphosphate (p4A) and diadenosine 5’,5’-tri-, tetra-, and pentaphosphates (Ap3A, Ap4A, Ap5A).87 The authors proposed that the enzyme reacts with ATP to form an activated AMP complex. Subsequent treatment of this complex with triphosphate, ADP, ATP, or p4A results respectively in the previously mentioned products. In similar reactions, Miller found that the stress protein LysU acts on ATP to form Ap4A,88 and Marx found that the ubiquitin activating enzyme UBA1 forms Ap3A and Ap4A.89 Furthermore, introduction of a mixture of ATP and other NTPs resulted in unsymmetric Ap4N analogues. Unsymmetric Ap4N species have also been synthesized using luciferase as catalyst.90 In an example that uses a genetically modified enzyme, Frey developed a modified human fragile histidine triad protein capable of catalyzing the reaction of NMP derived phosphorimidazolides and nucleotide substrates to make dinucleoside oligophosphates.91

Recent efforts have also shown potential utility for PPK enzymes that catalyze phosphoryl transfer from and to polyP. Jessen has shown that PPK2 not only catalyzes the transfer of one phosphoryl unit from polyP to ADP but can also competently utilize longer nucleotide substrates.92 In this manner, p4A and p5A were synthesized, with decreasing reactivity for longer nucleotides. Jendrossek then found that PPK2 from Agrobacterium tumefaciens is similarly capable of producing longer nucleotides with a greater distribution than the enzyme studied by Jessen, providing up to nucleoside nonaphosphates on an analytical scale.93 PolyP binding to PPK2 has also been analyzed crystallographically, providing further insight into the mechanism of these reactions.94

Biochemical Utilization of Oligophosphates

Oligophosphates are implicated in a variety of biochemical processes, resulting in use in assays and drug development. The most common oligophosphates are 5’-nucleotides, and p4A and p5A have been identified in a variety of biological systems. However, the full physiological impact of these extended nucleotides is not well known.2022 Nevertheless, several biological effects of p4A have been identified such as reduction in rabbit intraocular pressure95 and inducing vasodilation.96 Adenosine 5’-pentaphosphate (p5A) and uridine 5-pentaphosphate (p5U) have also been shown to strongly inhibit Ribonuclease A, and these nucleotides may have similar effects on a myriad of enzymes.97 Additional research may further elaborate the roles of extended nucleotides, but synthetic derivatives have found use in biochemical experiments as well. Most notably, nucleoside oligophosphates that have been terminally modified with fluorescent labels have been utilized in real time DNA sequencing.98,99 The longer oligophosphates result in polymerase activity comparable to that of free NTPs, as opposed to the much slower reactivity of terminally modified NTPs.

NpnNs have been relatively well studied, and NpnNs as long as heptaphosphates have been identified in biological systems.23,24 Physiologically, NpnNs are potent agonists of many P2Y receptors, and these compounds have been investigated with respect to blood clotting.43,102 For example, nucleotide pyrophosphatase/phosphodiesterase 4 (NPP4) was found to hydrolyze extracellular Ap3A and Ap4A, promoting platelet aggregation, and the authors suggested that development of inhibitors of this reaction may serve as antithrombotic drugs.103 Uridine adenosine tetraphosphate (Up4A) is also a vasoconstrictive agent, and it has been suggested as a potential treatment for cardiovascular disease.23 A number of studies have also found NpnNs to inhibit other enzymes such as adenylate kinase45 and ribonucleotide reductase.104 Denufosol, a Np4N, has even been clinically tested as a treatment for cystic fibrosis, albeit with mixed results.25 Another important area of study for Np4Ns is in HIV drug resistance. The nucleotide analogs commonly used in the treatment of HIV act by terminating growing viral DNA when incorporated, due to absence of a 3’ hydroxyl group. Resistance to these drugs is conferred via a reaction in which the nucleotide analog is cleaved from the DNA terminus by ATP, generating adenosine nucleoside tetraphosphates (Ap4N) and allowing reverse transcription to continue (Figure 6).105,106 Research has shown that other Np4Ns can be synthesized to inhibit this activity and potentially serve as drugs themselves via the reverse reaction.107

Figure 6.

Figure 6.

Simplified mechanism of HIV resistance to azidothymidine.

PolyP is ubiquitous in life and plays a variety of roles such as phosphate storage and part of blood clotting.26,84 Biosynthesis of polyP has been studied in order to use microorganisms to recover phosphate from wastewater, a process termed Enhanced Biological Phosphorus Removal (EBPR).108 However, research into the utilization of synthetically modified polyphosphates is scarcer. Several groups have sought to investigate proteins that bind polyP using biotinylated polyP. Jessen and Saiardi prepared a biotinylated octaphosphate to identify proteins in the human PolyPome,109 and utilization of longer functionalized PolyP has been made possible through EDAC mediated coupling developed by Morrissey.110 Several groups have utilized biotinylated PolyP containing an average of 700 phosphate units per molecule to similarly screen or immobilize PolyP binding proteins,110112 and this material has been made commercially available through Kerafast. The same EDAC mediated reaction has also been used to modify PolyP with fluorophores to visualize PolyP transport into cells.34 Another exciting development is the recent elucidation of protein polyphosphorylation as a nonenzymatic post translational modification, but it remains to be seen what new chemistry arises from investigation of this phenomenon.28

Synthesis of Extended Oligophosphates

Monophosphorylation

We define monophosphorylation as a reaction in which an electrophilic monophosphate synthon reacts with a nucleophile. Therefore, treatment of such a monophosphorylation reagent with a phosphate substrate results in an oligophosphate product bearing one additional phosphate group. A number of syntheses of oligophosphates have been reported using this approach, as many of the monophosphorylation reagents are the same as those that have been utilized for the synthesis of shorter oligophosphates, such as NTPs.

A large class of monophosphorylation reagents is based on electrophilic activation of an NMP. In a representative example, treatment of an NMP with CDI results in a phosphorimidazolide species.40 This imidazole group is a moderate strength leaving group that can be substituted by phosphate nucleophiles. A variety of additives have been used to accelerate these phosphorylation reactions and improve yields, commonly divalent metal salts such as MgCl2 and ZnCl2 as well as organic heterocycles such as 4,5-dicyanoimidazole (DCI).100 Many reagents other than CDI have also been used to similarly activate NMPs, such as DCC,18 acylating agents,113 sulfonyl reagents,,114 and 2-imidazolyl-1,3-dimethylimidazolinium chloride.33 Comparable phosphoramidate species have also been generated with various amine leaving groups such as morpholine115 and piperidine.100 This strategy has been utilized extensively with NMP and NDP substrates to synthesize dinucleoside di-and triphosphates,40,43,44,116 and extension of this methodology to longer oligophosphate substrates has given rise to a number of syntheses of extended oligophosphates.43,44,100,104 Rideout demonstrated synthesis of Up4U through several routes including treatment of UMP derived phosphorimidazolide with pyrophosphate (Figure 7).44 However, this synthesis provided only trace product contaminated by p4U, demonstrating the inefficiencies of these reactions in the absence of a suitable promoter. In a more successful example, Wang synthesized p4Ns in good yield from NMP derived phosphopiperidates utilizing DCI as a promoter (Figure 7).100 Furthermore, the isolated p4Ns were further monophosphorylated to produce Np5Ns in moderate yield (Figure 7). Another approach to increase the reactivity of monophosphorylation reagents is to synthesize zwitterionic species with tertiary amine derived leaving groups, such as pyridine, but this strategy has so far only been applied to the synthesis of shorter oligophosphates.117

Figure 7.

Figure 7.

Representative monophosphorylation syntheses of oligophosphates by Rideout,44 Wang,100 Kowalska,70 and Shanmugasundaram.101

Monophosphorylation reagents are not limited to those derived from NMPs, and the above strategy can be applied to activation of other organomonophosphates.70,98 However, electrophilic activation of inorganic orthophosphate presents a complication due to oligomerization.41 Accordingly, Kowalska reported a 2-cyanoethyl phosphorimidazolide, utilizing a protecting group widely used in oligonucleotide synthesis that is cleaved by mild base (Figure 7).70 This reagent reacts readily with NDP and NTP substrates in a microwave assisted reaction to extend the phosphate chain by one unit. Notably this methodology also imparts a facile way to incorporate thiophosphates into the oligophosphate, a common strategy with shorter oligophosphates that has not been well developed for longer derivatives.

Phosphochloridates are also competent monophosphorylation intermediates. Treatment of nucleosides with phosphoryl chloride in alkyl phosphate solvents, the Yoshikawa method, results in 5’-phosphochloridate-nucleosides.55 In the Ludwig synthesis of nucleoside triphosphates, these phosphochloridates are treated with pyrophosphate to generate a nucleoside-trimetaphosphate intermediate that is quenched with water to yield an NTP.118 In a comparable synthesis, Shanmugasundaram treated 5’-phosphochloridate-nucleosides with inorganic triphosphate, yielding p4Ns in 40–48% yield (Figure 7).101

Jessen has also demonstrated a variety of P(III) based monophosphorylation reactions. A fluorenylmethyl protected phosphoramidite was used to selectively and iteratively extend nucleotides through a monophosphorylation reaction, providing syntheses of NDPs, NTPs, and p4A.71,72 The same methodology has been extended to a wider variety of substrates, providing access to thiophosphates, sugarnucleotide conjugates, and Ap4U.119 Similarly, utilizing a phosphordiamidite reagent to couple two oligophosphate substrates generated products as long as Ap7A.120

Diphosphorylation

Diphosphorylation chemistry has not been as well explored as monophosphorylation, and these reactions are mostly limited to the synthesis of Np4Ns. The reaction of an activating agent at only one terminal phosphate of pyrophosphate would be expected to result in oligomerization. However, treating pyrophosphate with an excess of activating agent may result in a pyrophosphate species bearing a leaving group at both ends. This chemistry is best represented by the work of Wright, in which a pyrophosphate salt is treated with carbonyldiimidazole (CDI) to form a bis-phosphorimidazolide (Figure 8).79 This reagent reacts slowly with nucleoside monophosphates, a reaction that is accelerated with promoters such as ZnCl2, to give symmetric Np4Ns. This chemistry has not been demonstrated with longer substrates, such as nucleoside diphosphates, although this may be a route to dinucleoside hexaphosphates. Interestingly, the authors were able to synthesize similar diphosphorylation reagents from pyrophosphate analogs in which the bridging oxygen has been replaced with methylene derivatives, providing access to oligophosphates with nonhydrolyzable linkers.

Figure 8.

Figure 8.

Representative Diphosphorylation syntheses of Np4Ns.44,62,79 Thermal ellipsoid plot of crystallographically characterized diphosphorylation reagent P2O5(DABCO)2 with thermal ellipsoids set at 50% probability (red for oxygen, orange for phosphorus, blue for nitrogen, grey for carbon); this graphic was adapted from reference 62 as permitted under its CC BY 4.0 License https://creativecommons.org/licenses/by/4.0/.

A notable shortcoming of Wright’s diphosphorylation reagent is the difficulty of synthesizing unsymmetric Np4Ns. Recently, syntheses of neutral, dizwitterionic adducts of P2O5 with N-donor bases such as DABCO and pyridine have been reported (Figure 8).62 These neutral N-donor bases are superior leaving groups as compared with the anionic phosphorimidazolides used by Wright. In a stark contrast in reactivity to Wright’s reagent, the pyridine adduct of P2O5 reacts instantly with AMP to give an adenosine substituted TriMP intermediate. Ring opening of this intermediate with UMP generates unsymmetric Up4A in 75% yield. The extension of this methodology to longer oligophosphates similarly remains to be elucidated.

Diphosphorylation can also proceed through activation of a diphosphate compound. Similar to activation of NMPs, treatment of NDPs with CDI forms a terminal phosphorimidazolide, and this chemistry has been utilized to make symmetric Np4Ns.44,102 This methodology has so far suffered from low yields, and has only been utilized for the synthesis of symmetric Np4Ns. In principle, it should be possible to treat an activated NDP with a different NDP or with inorganic pyrophosphate to synthesize unsymmetric Np4Ns or p4Ns, but this chemistry has not been developed. Meier has also activated diphosphate compounds with trifluoroacetic anhydride. However, this methodology has only been used in the synthesis of triphosphate derivatives.122,123

Triphosphorylation

Activation of triphosphates results in an intramolecular dehydration to form cyclic TriMP derivatives. Therefore, triphosphorylation reactions proceed through a substituted TriMP intermediate, either through activation of a triphosphate substrate or directly from an activated TriMP reagent.

The traditional syntheses of NTPs by Khorana18 and Ludwig and Eckstein60 proceed through nucleoside substituted TriMP intermediates. Aqueous hydrolysis of these compounds yields NTPs, but ring opening with other nucleophiles is possible to give a wide variety of terminally functionalized NTPs.37,38 Ring opening can also be performed with phosphate nucleophiles to generate longer oligophosphates.35,37 However, ring opening with a phosphate nucleophile generally requires a promoter such as a divalent metal ion. In one example, adenosine substituted TriMP formed through Ludwig-Eckstein methodology was ring opened with AMP and ADP to provide Ap4A and Ap5A (Jones, Figure 9).121 A prime drawback of starting with Ludwig-Eckstein chemistry is the need for protection of the 2’ and 3’ hydroxyl groups. However, it is also possible to generate nucleoside substituted TriMP intermediates directly from isolated NTPs by activation, for example with DCC. Compounds generated in this manner have then been ring opened with NMPs to make Np4Ns43,44 or with NDPs to make Np5Ns.45

Figure 9.

Figure 9.

Representative triphosphorylation syntheses of oligophosphates by Jones,121 Taylor,35 and Jessen.37

In recent years, several TriMP based reagents have been developed to directly triphosphorylate a nucleophile rather than starting with an NTP substrate. Initial investigations attempted direct reactions of TriMP with NMPs, with some success generating p4Ns.124 However, this methodology produces a distribution of nucleoside oligophosphates, hampering its utility. The first electrophilic activation of TriMP comes from Taylor, who in situ activated TriMP with mesitylenesulfonyl chloride and nucleophilic tertiary amines,35,59,65,66 a reaction which has also been utilized by Kool.125 TriMP has been activated similarly with (7-azabenzotriazol-1-yloxy)tripyrrolidinophosphonium hexafluorophosphate, (PyAOP), and in this case the resulting activated TriMP derivative could be structurally characterized.61 Furthermore, it was reported subsequently that a derivative of TriMP activated with a sulfonyl chloride in direct analogy to Taylor’s procedure could be isolated and structurally characterized.62 Jessen37,38 and Huang67 have taken a related approach of synthesizing mixed P(III)-P(V) TriMP analogs that similarly react with nucleophiles to give substituted TriMP species. In one example, Taylor synthesized p4Ns and Np4Ns by treatment of activated TriMP with an NMP and ring opening of the resulting substituted metaphosphate either with water or another NMP (Figure 9) As a P(III) species, Jessen’s triphosphorylation reagent is highly reactive towards even difficult substrates such as NTPs. In addition to a number of other triphosphorylation reactions, Jessen has demonstrated triphosphorylation of ATP and ring opening of the subsequent TriMP intermediate to make a rare p6N derivative (Figure 9).37

Tetraphosphorylation

One of the few examples of a tetraphosphorylation reagent is [PPN]2[P4O11], the anhydride of dihydrogen tetrametaphosphate (TetraMP, Figure 10).126 This isolable and crystallographically characterized reagent has been treated with nucleosides and NMPs to make substituted TetraMP intermediates. Subsequent ring opening with hydroxide yielded p4Ns and p5Ns, which were studied as inhibitors of RNase A (Figure 10, top).97 In a follow up report, [PPN]2[P4O11] was further shown to be a competent tetraphosphorylation reagent for a variety of other nucleophiles and ring opening reactions, yielding a diverse class of tetra- and pentaphosphate derivatives, including an unusual example of a tetraphosphorylated amino acid, tyrosine.46

Figure 10.

Figure 10.

A. Tetraphosphorylation synthesis of p4Ns.97 B. Synthesis of a phosphonate analog of a 3’,4’ didehydronucleoside.46 C. Thermal ellipsoid plot of tetraphosphorylation reagent ([PPN]2[P4O11]) with thermal ellipsoids set at 50% probability and cations omitted for clarity (red for oxygen, orange for phosphorus).126 D. Tetraphosphorylation reaction utilizing P(III) oligophosphites by Parang.127

In an unusual example of C-phosphorylation, a previously reported TriMP based triphosphorylation reagent was used to demonstrate triphosphorylation of the phosphorus ylide methylenetriphenylphosphorane.61 Similarly, treatment of [PPN]2[P4O11] with methylenetriphenylphosphorane results in C–P bond formation followed by deprotonation to form a new phosphorus ylide (Figure 10, bottom).46 This ylide further reacts with aldehydes through Wittig chemistry and is therefore capable of generating tetraphosphate analogs where the terminal P–O bond has been replaced with a non-hydrolyzable olefinic P–C bond. In a particularly interesting example, treatment of the tetraphosphorylated ylide with a protected uridine derived aldehyde results in a tetraphosphate/phosphonate of a 3’,4’-didehydronucleoside (Figure 10, bottom).

Some tetraphosphorylation chemistry has also been developed with P(III) chemistry. In an unusual strategy, Parang developed a methodology to prepare oligophosphites, P(III) analogs of oligophosphates. Via iterative coupling and hydrolysis reactions of P(III) reagents, di-, tri-, and tetraphosphites were synthesized. These oligophosphites were then bound on a polymer support and treated with nucleosides. Subsequent oxidation and deprotection yielded symmetric Np2Ns, Np3Ns, and Np4Ns.127

Pentaphosphorylation and Beyond

A variety of monophosphorylation strategies exists, but the prevalence of methodologies decreases steeply for diphosphorylation and beyond. This is potentially explained by the comparatively rarer known applications for extended oligophosphates, the difficulty of obtaining larger phosphate starting materials, and the more difficult purification of extended oligophosphates. Nevertheless, there is a nascent field of phosphorylation reagents based on tri-35,37,38,59,61,65,66 and tetrametaphosphate.46,97,126 Larger metaphosphates, up to at least dodecametaphosphate, are fairly easily synthesized via flux methods,12 and we can naturally expect related chemistry derived from larger metaphosphates in the future. This may take the form of activation of these metaphosphates, similar to the work of Taylor35,59,65,66 and Cummins,46,61,97,126 or mixed P(III)-P(V) reagents like those of Jessen37,38,128 and Huang (Figure 11).67 Achieving high selectivities and yields with these larger reagents may prove difficult, but these routes seem promising for rational synthesis of well defined oligophosphate chain lengths.

Figure 11.

Figure 11.

Proposed penta- and hexaphosphorylation reagents in analogy to the crystalline tetraphosphorylation reagent [PPN]2[P4O11]46,97,126 and proposed mechanism of EDAC mediated functionalization of PolyP.

The scant methods for polyphosphorylation are not typically selective for a specific chain length. Rather, distributions of polyP starting materials are used, giving rise to a similar distribution of difficult to separate products. A variety of formulations of polyP are commercially available, often denoted as “hexametaphosphate” or Graham’s salt, usually produced by thermal condensation of lower phosphates.129,130 The distribution of phosphate chain lengths varies considerably and can be roughly analyzed by phosphorus NMR integration or SDS-PAGE.131 More precise quantifications of these distributions, perhaps via MALDI-TOF or capillary electrophoresis, remain desirable. Depending on the synthesis and supplier, these polyP formulations range in average phosphate chain length from P12 to >P500 and often contain large quantities of metaphosphates and short oligophosphates.

The polyP chemistry developed thus far has been aqueous. Aqueous systems prevent the use of water reactive species commonly used in phosphorylation chemistry, but aqueous media also provides the opportunity to phosphorylate biomolecules such as peptides and unprotected amino acids that are only appreciably soluble in water. Most of the known polyphosphorylation chemistry is based on Morrissey’s activation of polyphosphate with aqueous EDAC.110 The initial report proposes activation of the terminal phosphate units of PolyP to generate isourea leaving groups, which are intercepted by amine substrates to form terminally phosphoramidate modified polyphosphate.110,132 However, O-phosphoryl-isourea species are known to rapidly isomerize to N-phosphoryl-urea species which do not act as strong leaving groups.41 Therefore, we propose a modified mechanism involving EDAC mediated intramolecular dehydration (Figure 11). The resulting ultraphosphate species would then go on to react with nucleophilic substrates resulting in the final functionalized polyP. Also noteworthy is the fact that this functionalization is expected to occur at both ends of the polyP. This EDAC promoted polyphosphorylation procedure has since been expanded towards alcoholic substrates133 and been used in sporadic biochemical studies.34,111,112 Although activation of p3Ns has been widely used in triphosphorylation protocols,4345 a similar activation strategy has not been applied to well defined extended oligophosphates. For example, treatment of p4Ns or p5Ns with EDAC may provide an effective means of tetra- and pentaphosphorylation.

Polyphosphorylation of peptides has also recently grown in interest as an analog of protein monophosphorylation, the most common post translational modification.134 The last several decades have seen debate and investigation of how protein pyrophosphorylation fits into the canonical understanding of protein phosphorylation,135 and in 2015 Saiardi added the discovery of protein polyphosphorylation to this discussion.28 Polyphosphorylation was found to occur on lysine residues via a nonenzymatic reaction, and the authors went on to develop a yeast model to further study this post translational modification.136 The mechanism, importance, and pervasiveness of protein polyphosphorylation all remain areas of necessary study.

Outlook

The chemical synthesis of oligophosphates is simultaneously complex and nascent. Despite being a relatively small class of molecules, hundred of papers and multiple reviews detail the syntheses of nucleoside mono-, di-, and triphosphates.17,19,54 Whether due to difficulty of their synthesis or less demand for their applications, the synthesis of extended oligophosphates has certainly lagged behind that of their shorter analogs.

Further advances in the synthesis of pnNs and NpnNs are still possible to ease synthesis and purification, but a fair number of preparations of these compounds has been reported. Future synthetic work should endeavor to push past these compounds to more complex and varied products. A prime future direction should be the investigation of 3’-nucleotides. The biorelevance of such compounds is less well developed, but 3’-nucleoside oligophosphates have potential to interact with enzymes that bind DNA/RNA due to the 3’ linkages in these biopolymers. The synthesis of 3’-nucleotides also represents a somewhat greater challenge than analogous 5’-nucleotides due to the lower steric accessibility of this site, and the potential interference or hydrolysis promoted by the 2’ hydroxyl group. Another area for further exploration is nucleotide analogs derived from non ribose sugars, such as TNA triphosphates, which have been synthesized by Chaput.73,137 Dinucleoside oligophosphates similarly suffer from a dearth of 3’,3’ or mixed 3’,5’ species. One could even envision a cyclic product in which the 3’ and 5’ positions of a single nucleoside are bridged by an oligophosphate. Such cyclic pyrophosphate derivatives are reminiscent of pyrophosphate linked DNA138 and cyclic dinucleotides.139

Greater modification of the oligophosphate itself is also desirable. A significant class of ATP analogs are those in which the final bridging oxygen of the oligophosphate has been replaced with a non-hydrolyzable group, such as a methylene.140 This blocks hydrolysis to ADP, inhibiting enzymes and allowing biochemical study of their mechanisms. Such modifications are fairly common with shorter oligophosphates38 but are limited to a few examples for longer phosphates.79 Additionally, formation of thiophosphate analogs have not been reported for many large oligophosphates,70 despite thiophosphates being the basis for many drugs.141

A significant recent development in oligophosphorylation is the introduction of well-defined isolable oligophosphorylation reagents. The earliest oligophosphorylation procedures, such as Khorana’s synthesis of ATP,18 rely on in situ activation of nucleotides and inorganic phosphates. In the case of Khorana’s synthesis, the favorable formation of TriMPs results in a viable NTP synthesis, but the prevalence of side products results in low yield and difficult purification. In contrast, researchers have now developed bottleable reagents to append oligophosphates to a desired substrate.37,38,46,61,97,126,128 By introducing an activated phosphate reagent to a substrate of choice, this strategy reduces the potential side reactivity caused by activation of a phosphate substrate. For example, CDI activation of NMPs is an effective procedure,40 but there is still potential to form a 3’,5’-cyclic NMP side product. Utilizing a pre-synthesized (and potentially commercially available) reagent therefore simplifies the procedure and mitigates side reactivity.

Lastly, this Perspective has focused heavily on nucleosideoligophosphate conjugates and biochemical applications of these compounds, as this area has been the most developed. However, nucleotides are only a subclass of oligophosphorylated organic molecules, and there is potential for these compounds in a variety of fields of study and industry. For example geranyl diphosphate is a critical intermediate in terpene synthesis,142 biotinylated oligophosphates have been used in a number of studies,109112 and neither system contains nucleoside moieties. A much less developed but promising area is the utilization of oligophosphates in materials. A number of porous materials have been developed from zirconium phosphates and phosphonates,143145 but the utilization of oligophosphates for such applications has not to our knowledge been explored. Similarly, many phosphate and phosphonate containing polymers have been developed for biomedical or ion exchange purposes,146 but oligophosphates remain essentially unexplored. However, the field of chemical oligophosphorylation has flourished, providing a variety of simple syntheses of not only di- and triphosphates but extended oligophosphates as well. By introducing this chemistry to a wider audience, we hope this Perspective will stimulate further interest and study of extended oligophosphates, in the biochemical field and potentially far beyond.

Acknowledgement

C.C.C. acknowledges funding by the National Institutes of Health under award number R01GM130936. H.J.J. acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy - EXC-2193/1-390951807.

References

  • (1).Nicolaou K Organic synthesis: the art and science of replicating the molecules of living nature and creating others like them in the laboratory. Proc. Math. Phys. Eng. Sci 2014, 470, 20130690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (2).Amblard M; Fehrentz J-A; Martinez J; Subra G Methods and protocols of modern solid phase peptide synthesis. Mol. Biotechnol 2006, 33, 239–254. [DOI] [PubMed] [Google Scholar]
  • (3).Merrifield RB Solid phase peptide synthesis. I. The synthesis of a tetrapeptide. J. Am. Chem. Soc 1963, 85, 2149–2154. [Google Scholar]
  • (4).Fields GB; Noble RL Solid phase peptide synthesis utilizing 9-fluorenylmethoxycarbonyl amino acids. Int. J. Pept. Protein Res 1990, 35, 161–214. [DOI] [PubMed] [Google Scholar]
  • (5).Amarnath V; Broom A Chemical synthesis of oligonucleotides. Chem. Rev 1977, 77, 183–217. [Google Scholar]
  • (6).Sinha N; Biernat J; McManus J; Köster H Polymer support oligonucleotide synthesis XVIII1. 2): use of β-cyanoethyi-N, N-dialkylamino-/N-morpholino phosphoramidite of deoxynucleosides for the synthesis of DNA fragments simplifying deprotection and isolation of the final product. Nucleic Acids Res 1984, 12, 4539–4557. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Caruthers M; Barone A; Beaucage S; Dodds D; Fisher E; McBride L; Matteucci M; Stabinsky Z; Tang J-Y [15] Chemical synthesis of deoxyoligonucleotides by the phosphoramidite method. Methods Enzymol 1987, 154, 287–313. [DOI] [PubMed] [Google Scholar]
  • (8).Nicolaou K; Mitchell HJ Adventures in carbohydrate chemistry: new synthetic technologies, chemical synthesis, molecular design, and chemical biology. Angew. Chem., Int. Ed 2001, 40, 1576–1624. [PubMed] [Google Scholar]
  • (9).Kiessling LL; Splain RA Chemical approaches to glycobiology. Annu. Rev. Biochem 2010, 79, 619–653. [DOI] [PubMed] [Google Scholar]
  • (10).Elser J; Bennett E A broken biogeochemical cycle. Nature 2011, 478, 29–31. [DOI] [PubMed] [Google Scholar]
  • (11).White AK; Metcalf WW Microbial metabolism of reduced phosphorus compounds. Annu. Rev. Microbiol 2007, 61, 379–400. [DOI] [PubMed] [Google Scholar]
  • (12).Durif A Crystal chemistry of condensed phosphates; Springer Science & Business Media, 2013. [Google Scholar]
  • (13).Knowles JR Enzyme-catalyzed phosphoryl transfer reactions. Annu. Rev. Biochem 1980, 49, 877–919. [DOI] [PubMed] [Google Scholar]
  • (14).Sawaya MR; Pelletier H; Kumar A; Wilson SH; Kraut J Crystal structure of rat DNA polymerase beta: evidence for a common polymerase mechanism. Science 1994, 264, 1930–1935. [DOI] [PubMed] [Google Scholar]
  • (15).Qin K; Dong C; Wu G; Lambert NA Inactive-state preassembly of G q-coupled receptors and G q heterotrimers. Nature Chem. Biol 2011, 7, 740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Vaghefi M Nucleoside triphosphates and their analogs: chemistry, biotechnology, and biological applications; CRC press, 2016; pp 1–383. [Google Scholar]
  • (17).Roy B; Depaix A; Perigaud C; Peyrottes S Recent trends in nucleotide synthesis. Chem. Rev 2016, 116, 7854–7897. [DOI] [PubMed] [Google Scholar]
  • (18).Khorana H Carbodiimides. Part V. 1 A Novel Synthesis of Adenosine Di-and Triphosphate and P1, P2-Diadenosine-5’-pyrophosphate. J. Am. Chem. Soc 1954, 76, 3517–3522. [Google Scholar]
  • (19).Burgess K; Cook D Syntheses of nucleoside triphosphates. Chem. Rev 2000, 100, 2047–2060. [DOI] [PubMed] [Google Scholar]
  • (20).Small GD; Cooper C Studies on the occurrence and biosynthesis of adenosine tetraphosphate. Biochemistry 1966, 5, 26–33. [DOI] [PubMed] [Google Scholar]
  • (21).Guranowski A; Sillero MG; Sillero A Adenosine 5’-tetraphosphate and adenosine 5’-pentaphosphate are synthesized by yeast acetyl coenzyme A synthetase. J. Bacteriol 1994, 176, 2986–2990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Jakubowski H Sporulation of the yeast Saccharomyces cerevisiae is accompanied by synthesis of adenosine 5’-tetraphosphate and adenosine 5’-pentaphosphate. Proc. Natl. Acad. Sci. U.S.A 1986, 83, 2378–2382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (23).Jankowski V; Tölle M; Vanholder R; Schönfelder G; van der Giet M; Henning L; Schlüter H; Paul M; Zidek W; Jankowski J Uridine adenosine tetraphosphate: a novel endothelium-derived vasoconstrictive factor. Nat. Med 2005, 11, 223–227. [DOI] [PubMed] [Google Scholar]
  • (24).Jankowski J; Jankowski V; Laufer U; van der Giet M; Henning L; Tepel M; Zidek W; Schlüter H Identification and quantification of diadenosine polyphosphate concentrations in human plasma. Arterioscler., Thromb., Vasc. Biol 2003, 23, 1231–1238. [DOI] [PubMed] [Google Scholar]
  • (25).Kellerman D; Mospan AR; Engels J; Schaberg A; Gorden J; Smiley L Denufosol: a review of studies with inhaled P2Y2 agonists that led to Phase 3. Pulm. Pharmacol. Ther 2008, 21, 600–607. [DOI] [PubMed] [Google Scholar]
  • (26).Harold FM Inorganic polyphosphates in biology: structure, metabolism, and function. Bacteriol. Rev 1966, 30, 772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).Bhandari R; Saiardi A; Ahmadibeni Y; Snowman AM; Resnick AC; Kristiansen TZ; Molina H; Pandey A; Werner JK; Juluri KR, et al. Protein pyrophosphorylation by inositol pyrophosphates is a posttranslational event. Proc. Natl. Acad. Sci. U.S.A 2007, 104, 15305–15310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (28).Azevedo C; Livermore T; Saiardi A Protein polyphosphorylation of lysine residues by inorganic polyphosphate. Mol. Cell 2015, 58, 71–82. [DOI] [PubMed] [Google Scholar]
  • (29).Wu Y; Fa M; Tae EL; Schultz PG; Romesberg FE Enzymatic phosphorylation of unnatural nucleosides. J. Am. Chem. Soc 2002, 124, 14626–14630. [DOI] [PubMed] [Google Scholar]
  • (30).Wen L; Huang K; Liu Y; Wang PG Facile enzymatic synthesis of phosphorylated ketopentoses. ACS Catal 2016, 6, 1649–1654. [Google Scholar]
  • (31).Ni F; Sun S; Huang C; Zhao Y N-phosphorylation of amino acids by trimetaphosphate in aqueous solution–learning from prebiotic synthesis. Green Chem 2009, 11, 569–573. [Google Scholar]
  • (32).Marmelstein AM; Yates LM; Conway JH; Fiedler D Chemical pyrophosphorylation of functionally diverse peptides. J. Am. Chem. Soc 2014, 136, 108–111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (33).Tanaka H; Yoshimura Y; Jørgensen MR; Cuesta-Seijo JA; Hindsgaul O A simple synthesis of sugar nucleoside diphosphates by chemical coupling in water. Angew. Chem 2012, 124, 11699–11702. [DOI] [PubMed] [Google Scholar]
  • (34).Fernandes-Cunha GM; McKinlay CJ; Vargas JR; Jessen HJ; Waymouth RM; Wender PA Delivery of inorganic polyphosphate into cells using amphipathic oligocarbonate transporters. ACS Cent. Sci 2018, 4, 1394–1402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Mohamady S; Taylor SD Synthesis of nucleoside tetraphosphates and dinucleoside pentaphosphates via activation of cyclic trimetaphosphate. Org. Lett 2013, 15, 2612–2615. [DOI] [PubMed] [Google Scholar]
  • (36).Chakarawet K; Knopf I; Nava M; Jiang Y; Stauber JM; Cummins CC Crystalline metaphosphate acid salts: synthesis in organic media, structures, hydrogen-bonding capability, and implication of superacidity. Inorg. Chem 2016, 55, 6178–6185. [DOI] [PubMed] [Google Scholar]
  • (37).Singh J; Steck N; De D; Hofer A; Ripp A; Captain I; Keller M; Wender PA; Bhandari R; Jessen HJ A phosphoramidite analogue of cyclotriphosphate enables iterative polyphosphorylations. Angew. Chem., Int. Ed 2019, 58, 3928–3933. [DOI] [PubMed] [Google Scholar]
  • (38).Singh J; Ripp A; Haas TM; Qiu D; Keller M; Wender PA; Siegel JS; Baldridge KK; Jessen HJ Synthesis of modified nucleoside oligophosphates simplified: Fast, pure, and protecting group free. J. Am. Chem. Soc 2019, 141, 15013–15017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (39).Mohamady S; Jakeman DL An improved method for the synthesis of nucleoside triphosphate analogues. J. Org. Chem 2005, 70, 10588–10591. [DOI] [PubMed] [Google Scholar]
  • (40).Zatorski A; Goldstein BM; Colby TD; Jones JP; Pankiewicz KW Potent inhibitors of human inosine monophosphate dehydrogenase type II. fluorine-substituted analogs of thiazole-4-carboxamide adenine dinucleotide. J. Med. Chem 1995, 38, 1098–1105. [DOI] [PubMed] [Google Scholar]
  • (41).Glonek T; Van Wazer JR; Myers TC Carbodiimide-Mediated Esterifications and Condensations of Phosphoric Acids Dissolved in Various Alcohols. Phosphorus, Sulfur Silicon Relat. Elem 1977, 3, 137–150. [Google Scholar]
  • (42).Ng K.-m. E.; Orgel LE The action of a water-soluble carbodiimide on adenosine-5’-polyphosphates. Nucleic Acids Res 1987, 15, 3573–3580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (43).Shaver SR; Rideout JL; Pendergast W; Douglass JG; Brown EG; Boyer JL; Patel RI; Redick CC; Jones AC; Picher M, et al. Structure–activity relationships of dinucleotides: potent and selective agonists of P2Y receptors. Purinergic Signal 2005, 1, 183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (44).Pendergast W; Yerxa BR; Douglass III JG; Shaver SR; Dougherty RW; Redick CC; Sims IF; Rideout JL Synthesis and P2Y receptor activity of a series of uridine dinucleoside 5’-polyphosphates. Bioorg. Med. Chem. Lett 2001, 11, 157–160. [DOI] [PubMed] [Google Scholar]
  • (45).Hampton A; Kappler F; Picker D Species-or isozymespecific enzyme inhibitors. 4. Design of a two-site inhibitor of adenylate kinase with isozyme selectivity. J. Med. Chem 1982, 25, 638–644. [DOI] [PubMed] [Google Scholar]
  • (46).Shepard SM; Kim H; Bang QX; Alhokbany N; Cummins CC Synthesis of α, δ-Disubstituted Tetraphosphates and Terminally-Functionalized Nucleoside Pentaphosphates. J. Am. Chem. Soc 2020, 143, 463–470. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (47).Tsay F-R; Ahmad IAH; Henderson D; Schiavone N; Liu Z; Makarov AA; Mangion I; Regalado EL Generic anion-exchange chromatography method for analytical and preparative separation of nucleotides in the development and manufacture of drug substances. J. Chromatogr. A 2019, 1587, 129–135. [DOI] [PubMed] [Google Scholar]
  • (48).Perrone PA; Brown PR Ion-pair chromatography of nucleotides. J. Chromatogr. A 1984, 317, 301–310. [Google Scholar]
  • (49).Zhou T; Lucy CA Hydrophilic interaction chromatography of nucleotides and their pathway intermediates on titania. J. Chromatogr. A 2008, 1187, 87–93. [DOI] [PubMed] [Google Scholar]
  • (50).Geldart SE; Brown PR Analysis of nucleotides by capillary electrophoresis. J. Chromatogr. A 1998, 828, 317–336. [DOI] [PubMed] [Google Scholar]
  • (51).Qiu D; Wilson MS; Eisenbeis VB; Harmel RK; Riemer E; Haas TM; Wittwer C; Jork N; Gu C; Shears SB, et al. Analysis of inositol phosphate metabolism by capillary electrophoresis electrospray ionization mass spectrometry. Nat. Commun 2020, 11, 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (52).Blüher D; Laha D; Thieme S; Hofer A; Eschen-Lippold L; Masch A; Balcke G; Pavlovic I; Nagel O; Schonsky A, et al. A 1-phytase type III effector interferes with plant hormone signaling. Nat. Commun 2017, 8, 1–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (53).Corbridge DE Phosphorus: chemistry, biochemistry and technology; CRC press, 2013; pp 1–1437. [Google Scholar]
  • (54).Johnson DC; Widlanski TS Overview of the synthesis of nucleoside phosphates and polyphosphates. Curr. Protoc. Nucleic Acid Chem 2003, 15, 13–1. [DOI] [PubMed] [Google Scholar]
  • (55).Yoshikawa M; Kato T; Takenishi T A novel method for phosphorylation of nucleosides to 5’-nucleotides. Tetrahedron Lett 1967, 8, 5065–5068. [DOI] [PubMed] [Google Scholar]
  • (56).Knouse KW; Justine N; Schmidt MA; Zheng B; Vantourout JC; Kingston C; Mercer SE; Mcdonald IM; Olson RE; Zhu Y, et al. Unlocking P (V): Reagents for chiral phosphorothioate synthesis. Science 2018, 361, 1234–1238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (57).Huang Y; Knouse KW; Qiu S; Hao W; Padial NM; Vantourout JC; Zheng B; Mercer SE; Lopez-Ogalla J; Narayan R, et al. AP (V) platform for oligonucleotide synthesis. Science 2021, 373, 1265–1270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (58).Davisson VJ; Davis DR; Dixit VM; Poulter CD Synthesis of nucleotide 5’-diphosphates from 5’-O-tosyl nucleosides. J. Org. Chem 1987, 52, 1794–1801. [Google Scholar]
  • (59).Mohamady S; Taylor SD Synthesis of nucleoside triphosphates from 2’–3’-protected nucleosides using trimetaphosphate. Org. Lett 2016, 18, 580–583. [DOI] [PubMed] [Google Scholar]
  • (60).Ludwig J; Eckstein F Rapid and efficient synthesis of nucleoside 5’–0-(1-thiotriphosphates), 5’-triphosphates and 2’, 3’-cyclophosphorothioates using 2-chloro-4H-1, 3, 2-benzodioxaphosphorin-4-one. J. Org. Chem 1989, 54, 631–635. [Google Scholar]
  • (61).Shepard SM; Cummins CC Functionalization of intact trimetaphosphate: a triphosphorylating reagent for C, N, and O nucleophiles. J. Am. Chem. Soc 2019, 141, 1852–1856. [DOI] [PubMed] [Google Scholar]
  • (62).Shepard SM; Cummins CC N-donor Base Adducts of P2O5 as Diphosphorylation Reagents. ChemRxiv June 25, 2021, 10.26434/chemrxiv-2021-8v8nx. (date accessed 2021–12–12). [DOI] [Google Scholar]
  • (63).Engelsma SB; Meeuwenoord NJ; Overkleeft HS; van der Marel GA; Filippov DV Combined Phosphoramidite-Phosphodiester Reagents for the Synthesis of Methylene Bisphosphonates. Angew. Chem 2017, 129, 3001–3005. [DOI] [PubMed] [Google Scholar]
  • (64).Fiedler D; Haucke V; Türkaydin B; Furkert D; Dornan GL; Sun H; Utesch T; Hostachy S; Franke K Dissecting the activation of insulin degrading enzyme by inositol pyrophosphates and their bisphosphonate analogs. Chem. Sci 2021, 12, 10696–10702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (65).Mohamady S; Taylor SD Synthesis of nucleoside 5’-tetraphosphates containing terminal fluorescent labels via activated cyclic trimetaphosphate. J. Org. Chem 2014, 79, 2308–2313. [DOI] [PubMed] [Google Scholar]
  • (66).Mohamady S; Taylor SD Synthesis of Nucleoside-5’-O-Tetraphosphates from Activated Trimetaphosphate and Nucleoside-5’-O-Monophosphates. Curr. Protoc. Nucleic Acid Chem 2018, 75, e62. [DOI] [PubMed] [Google Scholar]
  • (67).Caton-Williams J; Hoxhaj R; Fiaz B; Huang Z Use of a Novel 5’-Regioselective Phosphitylating Reagent for One-Pot Synthesis of Nucleoside 5’-Triphosphates from Unprotected Nucleosides. Curr. Protoc. Nucleic Acid Chem 2013, 52, 1–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (68).Liao J-Y; Bala S; Ngor AK; Yik EJ; Chaput JC P (V) Reagents for the Scalable Synthesis of Natural and Modified Nucleoside Triphosphates. J. Am. Chem. Soc 2019, 141, 13286–13289. [DOI] [PubMed] [Google Scholar]
  • (69).Mohamady S; Taylor SD General procedure for the synthesis of dinucleoside polyphosphates. J. Org. Chem 2011, 76, 6344–6349. [DOI] [PubMed] [Google Scholar]
  • (70).Strenkowska M; Wanat P; Ziemniak M; Jemielity J; Kowalska J Preparation of synthetically challenging nucleotides using cyanoethyl P-imidazolides and microwaves. Org. Lett 2012, 14, 4782–4785. [DOI] [PubMed] [Google Scholar]
  • (71).Jessen HJ; Ahmed N; Hofer A Phosphate esters and anhydrides–recent strategies targeting nature’s favoured modifications. Org. Biomol. Chem 2014, 12, 3526–3530. [DOI] [PubMed] [Google Scholar]
  • (72).Cremosnik GS; Hofer A; Jessen HJ Iterative synthesis of nucleoside oligophosphates with phosphoramidites. Angew. Chem., Int. Ed 2014, 53, 286–289. [DOI] [PubMed] [Google Scholar]
  • (73).Sau SP; Chaput JC A gram-scale HPLC-free synthesis of TNA triphosphates using an iterative phosphorylation strategy. Org. Lett 2017, 19, 4379–4382. [DOI] [PubMed] [Google Scholar]
  • (74).Wu W; Bergstrom DE; Davisson VJ A combination chemical and enzymatic approach for the preparation of azole carboxamide nucleoside triphosphate. J. Org. Chem 2003, 68, 3860–3865. [DOI] [PubMed] [Google Scholar]
  • (75).Hennig M; Scott LG; Sperling E; Bermel W; Williamson JR Synthesis of 5-fluoropyrimidine nucleotides as sensitive NMR probes of RNA structure. J. Am. Chem. Soc 2007, 129, 14911–14921. [DOI] [PubMed] [Google Scholar]
  • (76).Williams R Phosphorus biochemistry. Phosphorus in the Environment: Its Chemistry and Biochemistry 1978, 5, 95–116. [Google Scholar]
  • (77).Subramani S; Jayapalan S; Kalpana R; Natarajan J HomoKinase: a curated database of human protein kinases. ISRN Comput. Biol 2013, 2013, 417634. [Google Scholar]
  • (78).Mildvan A Role of magnesium and other divalent cations in ATP-utilizing enzymes. Magnesium 1987, 6, 28–33. [PubMed] [Google Scholar]
  • (79).Yanachkov IB; Dix EJ; Yanachkova MI; Wright GE P1, P2-diimidazolyl derivatives of pyrophosphate and bisphosphonates–synthesis, properties, and use in preparation of dinucleoside tetraphosphates and analogs. Org. Biomol. Chem 2011, 9, 730–738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (80).Wood HG; Clark JE Biological aspects of inorganic polyphosphates. Annu. Rev. Biochem 1988, 57, 235–260. [DOI] [PubMed] [Google Scholar]
  • (81).Zhang H; Ishige K; Kornberg A A polyphosphate kinase (PPK2) widely conserved in bacteria. Proc. Natl. Acad. Sci. U. S. A 2002, 99, 16678–16683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (82).Mandala VS; Loh DM; Shepard SM; Geeson MB; Sergeyev IV; Nocera DG; Cummins CC; Hong M Bacterial Phosphate Granules Contain Cyclic Polyphosphates: Evidence from 31P Solid-State NMR. J. Am. Chem. Soc 2020, 142, 18407–18421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (83).Ruiz FA; Lea CR; Oldfield E; Docampo R Human platelet dense granules contain polyphosphate and are similar to acidocalcisomes of bacteria and unicellular eukaryotes. J. Biol. Chem 2004, 279, 44250–44257. [DOI] [PubMed] [Google Scholar]
  • (84).Morrissey JH; Choi SH; Smith SA Polyphosphate: an ancient molecule that links platelets, coagulation, and inflammation. Blood 2012, 119, 5972–5979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (85).Desfougères Y; Saiardi A; Azevedo C Inorganic polyphosphate in mammals: where’s Wally? Biochem. Soc. Trans 2020, 48, 95–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (86).Baev AY; Angelova PR; Abramov AY Inorganic polyphosphate is produced and hydrolyzed in F0F1-ATP synthase of mammalian mitochondria. Biochem. J 2020, 477, 1515–1524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (87).Nakajima H; Tomioka I; Kitabatake S; Dombou M; Tomita K Facile and selective synthesis of diadenosine polyphosphates through catalysis by leucyl t-RNA synthetase coupled with ATP regeneration. Agr. Biol. Chem 1989, 53, 615–623. [Google Scholar]
  • (88).Theoclitou M-E; Wittung EPL; Hindley AD; El-Thaher TS; Miller AD Characterisation of stress protein LysU. Enzymic synthesis of diadenosine 5’-, 5”’-P1,P4-tetraphosphate (Ap4A) analogues by LysU. J. Chem. Soc. Perkin Trans. I 1996, 2009–2019. [Google Scholar]
  • (89).Götz KH; Mex M; Stuber K; Offensperger F; Scheffner M; Marx A Formation of the alarmones diadenosine triphosphate and tetraphosphate by ubiquitin-and ubiquitin-like-activating enzymes. Cell Chem. Bio 2019, 26, 1535–1543. [DOI] [PubMed] [Google Scholar]
  • (90).Ortiz B; Sillero A; Günther Sillero MA Specific synthesis of adenosine (5’) tetraphospho (5’) nucleoside and adenosine (5’) oligophospho (5’) adenosine (n> 4) catalyzed by firefly luciverase. Eur. J. Biochem 1993, 212, 263–270. [DOI] [PubMed] [Google Scholar]
  • (91).Huang K; Frey PA Engineering human Fhit, a diadenosine triphosphate hydrolase, into an efficient dinucleoside polyphosphate synthase. J. Am. Chem. Soc 2004, 126, 9548–9549. [DOI] [PubMed] [Google Scholar]
  • (92).Mordhorst S; Singh J; Mohr MK; Hinkelmann R; Keppler M; Jessen HJ; Andexer JN Several Polyphosphate Kinase 2 Enzymes Catalyse the Production of Adenosine 5’-Polyphosphates. ChemBioChem 2019, 20, 1019–1022. [DOI] [PubMed] [Google Scholar]
  • (93).Frank C; Teleki A; Jendrossek D Characterization of Agrobacterium tumefaciens PPKs reveals the formation of oligophosphorylated products up to nucleoside nonaphosphates. Appl. Microbiol. Biotechnol 2020, 104, 9683–9692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (94).Parnell AE; Mordhorst S; Kemper F; Giurrandino M; Prince JP; Schwarzer NJ; Hofer A; Wohlwend D; Jessen HJ; Gerhardt S, et al. Substrate recognition and mechanism revealed by ligand-bound polyphosphate kinase 2 structures. Proc. Natl. Acad. Sci. U. S. A 2018, 115, 3350–3355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (95).Pintor J; Pelaez T; Peral A Adenosine tetraphosphate, Ap4, a physiological regulator of intraocular pressure in normotensive rabbit eyes. J. Pharmacol. Exp. Ther 2004, 308, 468–473. [DOI] [PubMed] [Google Scholar]
  • (96).Westhoff T; Jankowski J; Schmidt S; Luo J; Giebing G; Schlüter H; Tepel M; Zidek W; van der Giet M Identification and characterization of adenosine 5’-tetraphosphate in human myocardial tissue. J. Biol. Chem 2003, 278, 17735–17740. [DOI] [PubMed] [Google Scholar]
  • (97).Shepard SM; Windsor IW; Raines RT; Cummins CC Nucleoside tetra-and pentaphosphates prepared using a tetraphosphorylation reagent are potent inhibitors of ribonuclease A. J. Am. Chem. Soc 2019, 141, 18400–18404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (98).Sood A; Kumar S; Nampalli S; Nelson JR; Macklin J; Fuller CW Terminal phosphate-labeled nucleotides with improved substrate properties for homogeneous nucleic acid assays. J. Am. Chem. Soc 2005, 127, 2394–2395. [DOI] [PubMed] [Google Scholar]
  • (99).Eid J; Fehr A; Gray J; Luong K; Lyle J; Otto G; Peluso P; Rank D; Baybayan P; Bettman B, et al. Real-time DNA sequencing from single polymerase molecules. Science 2009, 323, 133–138. [DOI] [PubMed] [Google Scholar]
  • (100).Sun Q; Sun J; Gong S-S; Wang C-J; Wang X-C Synthesis of nucleoside tetraphosphates and dinucleoside pentaphosphates from nucleoside phosphoropiperidates via the activation of P(V)-N bond. Chin. Chem. Lett 2015, 26, 89–92. [Google Scholar]
  • (101).Kore AR; Senthilvelan A; Srinivasan B; Shanmugasundaram M Facile protection-free one-pot chemical synthesis of nucleoside-5’-tetraphosphates. Nucleosides, Nucleotides Nucleic Acids 2013, 32, 411–420. [DOI] [PubMed] [Google Scholar]
  • (102).Douglass JG; Patel RI; Yerxa BR; Shaver SR; Watson PS; Bednarski K; Plourde R; Redick CC; Brubaker K; Jones AC, et al. Lipophilic modifications to dinucleoside polyphosphates and nucleotides that confer antagonist properties at the platelet P2Y12 receptor. J. Med. Chem 2008, 51, 1007–1025. [DOI] [PubMed] [Google Scholar]
  • (103).Lopez V; Lee S-Y; Stephan H; Müller CE Recombinant expression of ecto-nucleotide pyrophosphatase/phosphodiesterase 4 (NPP4) and development of a luminescence-based assay to identify inhibitors. Anal. Biochem 2020, 603, 113774. [DOI] [PubMed] [Google Scholar]
  • (104).Davies LC; Stock JA; Barrie SE; Orr RM; Harrap KR Dinucleotide analogs as inhibitors of thymidine kinase, thymidylate kinase and ribonucleotide reductase. J. Med. Chem 1988, 31, 1305–1308. [DOI] [PubMed] [Google Scholar]
  • (105).Sarafianos SG; Hughes SH; Arnold E Designing anti-AIDS drugs targeting the major mechanism of HIV-1 RT resistance to nucleoside analog drugs. Int. J. Biochem. Cell Biol 2004, 36, 1706–1715. [DOI] [PubMed] [Google Scholar]
  • (106).Meyer PR; Matsuura SE; Mian AM; So AG; Scott WA A mechanism of AZT resistance: an increase in nucleotide-dependent primer unblocking by mutant HIV-1 reverse transcriptase. Mol. Cell 1999, 4, 35–43. [DOI] [PubMed] [Google Scholar]
  • (107).Meyer PR; Smith AJ; Matsuura SE; Scott WA Chain-terminating dinucleoside tetraphosphates are substrates for DNA polymerization by HIV-1 reverse transcriptase with increased activity against thymidine analogue resistant mutants. Antimicrob. Agents Chemother 2006, 50, 3607–3614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (108).Oehmen A; Lemos PC; Carvalho G; Yuan Z; Keller J; Blackall LL; Reis MA Advances in enhanced biological phosphorus removal: from micro to macro scale. Water Res 2007, 41, 2271–2300. [DOI] [PubMed] [Google Scholar]
  • (109).Azevedo C; Singh J; Steck N; Hofer A; Ruiz FA; Singh T; Jessen HJ; Saiardi A Screening a protein array with synthetic biotinylated inorganic polyphosphate to define the human polyP-ome. ACS Chem. Biol 2018, 13, 1958–1963. [DOI] [PubMed] [Google Scholar]
  • (110).Choi SH; Collins JN; Smith SA; Davis-Harrison RL; Rienstra CM; Morrissey JH Phosphoramidate end labeling of inorganic polyphosphates: facile manipulation of polyphosphate for investigating and modulating its biological activities. Biochem 2010, 49, 9935–9941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (111).Khong ML; Li L; Solesio ME; Pavlov EV; Tanner JA Inorganic polyphosphate controls cyclophilin B-mediated collagen folding in osteoblast-like cells. FEBS J 2020, 287, 4500–4524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (112).Smith SA; Choi SH; Collins JN; Travers RJ; Cooley BC; Morrissey JH Inhibition of polyphosphate as a novel strategy for preventing thrombosis and inflammation. Blood 2012, 120, 5103–5110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (113).Kim KB; Behrman EJ A New Synthesis of Symmetrical P1,P2-Dinucleoside-5’-pyrophosphates. Nucleosides Nucleotides 1999, 18, 51–54. [Google Scholar]
  • (114).Mohamady S; Desoky A; Taylor SD Sulfonyl imidazolium salts as reagents for the rapid and efficient synthesis of nucleoside polyphosphates and their conjugates. Org. Lett 2012, 14, 402–405. [DOI] [PubMed] [Google Scholar]
  • (115).Van Boom J; Crea R; Luyten W; Vink A 2, 2, 2-tribromoethyl phosphoromorphalinochlorioate: A convenient reagent for the synthesis of ribonucleoside mono-, di-and triphosphates. Tetrahedron Lett 1975, 16, 2779–2782. [Google Scholar]
  • (116).Stern N; Major DT; Gottlieb HE; Weizman D; Fischer B What is the conformation of physiologically-active dinucleoside polyphosphates in solution? Conformational analysis of free dinucleoside polyphosphates by NMR and molecular dynamics simulations. Org. Biomol. Chem 2010, 8, 4637–4652. [DOI] [PubMed] [Google Scholar]
  • (117).Sun Q; Liu S; Sun J; Gong S; Xiao Q; Shen L One-pot synthesis of symmetrical P1, P2-dinucleoside-5’-diphosphates from nucleoside-5’-H-phosphonates: mechanistic insights into reaction path. Tetrahedron Lett 2013, 54, 3842–3845. [Google Scholar]
  • (118).Ludwig J A new route to nucleoside 5’-triphosphates. Acta Biochim. Biophys. Acad. Sci. Hung 1981, 16, 131–133. [PubMed] [Google Scholar]
  • (119).Hofer A; Cremosnik GS; Müller AC; Giambruno R; Trefzer C; Superti-Furga G; Bennett KL; Jessen HJ A modular synthesis of modified phosphoanhydrides. Chem. -Eur. J 2015, 21, 10116–10122. [DOI] [PubMed] [Google Scholar]
  • (120).Hofer A; Marques E; Kieliger N; Gatter S-KN; Jordi S; Ferrari E; Hofmann M; Fitzpatrick TB; Hottiger MO; Jessen HJ Chemoselective dimerization of phosphates. Org. Lett 2016, 18, 3222–3225. [DOI] [PubMed] [Google Scholar]
  • (121).Han Q; Gaffney BL; Jones RA One-flask synthesis of dinucleoside tetra-and pentaphosphates. Org. Lett 2006, 8, 2075–2077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (122).Zhao C; Weber S; Schols D; Balzarini J; Meier C Prodrugs of γ-Alkyl-Modified Nucleoside Triphosphates: Improved Inhibition of HIV Reverse Transcriptase. Angew. Chem., Int. Ed 2020, 59, 22063–22071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (123).Gollnest T; Dinis de Oliveira T; Rath A; Hauber I; Schols D; Balzarini J; Meier C Membrane-permeable Triphosphate Prodrugs of Nucleoside Analogues. Angew. Chem., Int. Ed 2016, 55, 5255–5258. [DOI] [PubMed] [Google Scholar]
  • (124).Lohrmann R Formation of nucleoside 5’-polyphosphates from nucleotides and trimetaphosphate. J. Mol. Evol 1975, 6, 237–252. [DOI] [PubMed] [Google Scholar]
  • (125).Mohsen MG; Ji D; Kool ET Polymerase synthesis of four-base DNA from two stable dimeric nucleotides. Nucleic Acids Res 2019, 47, 9495–9501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (126).Jiang Y; Chakarawet K; Kohout AL; Nava M; Marino N; Cummins CC Dihydrogen Tetrametaphosphate,[P4O12H2] 2−: Synthesis, Solubilization in Organic Media, Preparation of its Anhydride [P4O11]2− and Acidic Methyl Ester, and Conversion to Tetrametaphosphate Metal Complexes via Protonolysis. J. Am. Chem. Soc 2014, 136, 11894–11897. [DOI] [PubMed] [Google Scholar]
  • (127).Ahmadibeni Y; Parang K Solid-Phase Synthesis of Symmetrical 5’,5’-Dinucleoside Mono-, Di-, Tri-, and Tetraphosphodiesters. Org. Lett 2007, 9, 4483–4486. [DOI] [PubMed] [Google Scholar]
  • (128).Haas TM; Qiu D; Häner M; Angebauer L; Ripp A; Singh J; Koch H-G; Jessen-Trefzer C; Jessen HJ Four phosphates at one blow: Access to pentaphosphorylated magic spot nucleotides and their analysis by capillary electrophoresis. J. Org. Chem 2020, 85, 14496–14506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (129).McCullough J; Van Wazer J; Griffth E Structure and properties of the condensed phosphates. XI. Hydrolytic degradation of Graham’s salt. J. Am. Chem. Soc 1956, 78, 4528–4533. [Google Scholar]
  • (130).Bhargava H; Srivastava D; Varma B Molecular-weight distribution in Graham’s salt: Dependence on conditions of formation. Colloid Polym. Sci 1974, 252, 20–25. [Google Scholar]
  • (131).Bondy-Chorney E; Abramchuk I; Nasser R; Holinier C; Denoncourt A; Baijal K; McCarthy L; Khacho M; Lavallée-Adam M; Downey M A broad response to intracellular long-chain polyphosphate in human cells. Cell Rep 2020, 33, 108318. [DOI] [PubMed] [Google Scholar]
  • (132).Baker CJ; Smith SA; Morrissey JH Diversification of polyphosphate end-labeling via bridging molecules. PLOS One 2020, 15, e0237849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (133).Hebbard CF; Wang Y; Baker CJ; Morrissey JH Synthesis and evaluation of chromogenic and fluorogenic substrates for high-throughput detection of enzymes that hydrolyze inorganic polyphosphate. Biomacromolecules 2014, 15, 3190–3196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (134).Cohen P The origins of protein phosphorylation. Nat. Cell Biol 2002, 4, E127–E130. [DOI] [PubMed] [Google Scholar]
  • (135).Saiardi A Protein pyrophosphorylation: moving forward. Biochem. J 2016, 473, 3765–3768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (136).Azevedo C; Desfougères Y; Jiramongkol Y; Partington H; Trakansuebkul S; Singh J; Steck N; Jessen HJ; Saiardi A Development of a yeast model to study the contribution of vacuolar polyphosphate metabolism to lysine polyphosphorylation. J. Biol. Chem 2020, 295, 1439–1451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (137).Bala S; Liao J-Y; Mei H; Chaput JC Synthesis of α-L-threofuranosyl nucleoside 3’-monophosphates, 3’-phosphoro (2-methyl) imidazolides, and 3’-triphosphates. J. Org. Chem 2017, 82, 5910–5916. [DOI] [PubMed] [Google Scholar]
  • (138).Kistemaker HA; Meeuwenoord NJ; Overkleeft HS; van der Marel GA; Filippov DV On the synthesis of oligonucleotides interconnected through pyrophosphate linkages. Eur. J. Org. Chem 2015, 2015, 6084–6091. [Google Scholar]
  • (139).Danilchanka O; Mekalanos JJ Cyclic dinucleotides and the innate immune response. Cell 2013, 154, 962–970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (140).Bystrom CE; Pettigrew DW; Remington SJ; Branchaud BP ATP analogs with non-transferable groups in the γ position as inhibitors of glycerol kinase. Bioorg. Med. Chem. Lett 1997, 7, 2613–2616. [Google Scholar]
  • (141).Gleave ME; Monia BP Antisense therapy for cancer. Nat. Rev. Cancer 2005, 5, 468–479. [DOI] [PubMed] [Google Scholar]
  • (142).Burke CC; Wildung MR; Croteau R Geranyl diphosphate synthase: cloning, expression, and characterization of this prenyltransferase as a heterodimer. Proc. Natl. Acad. Sci. U.S.A 1999, 96, 13062–13067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (143).Clearfield A; Wang Z Organically pillared microporous zirconium phosphonates. Dalton Trans 2002, 2937–2947. [Google Scholar]
  • (144).Silbernagel R; Martin CH; Clearfield A Zirconium (IV) phosphonate–phosphates as efficient ion-exchange materials. Inorg. Chem 2016, 55, 1651–1656. [DOI] [PubMed] [Google Scholar]
  • (145).Casciola M From layered zirconium phosphates and phosphonates to nanofillers for ionomeric membranes. Solid State Ionics 2019, 336, 1–10. [Google Scholar]
  • (146).Monge S; Canniccioni B; Graillot A; Robin J-J Phosphorus-containing polymers: a great opportunity for the biomedical field. Biomacromolecules 2011, 12, 1973–1982. [DOI] [PubMed] [Google Scholar]

RESOURCES