Fact-checked by Grok 2 weeks ago

Oligonucleotide synthesis

Oligonucleotide synthesis is the chemical process by which short, single-stranded nucleic acid sequences—such as DNA or RNA—are assembled nucleotide by nucleotide to produce custom oligonucleotides typically 10 to 200 bases in length. This automated technique enables the rapid production of these molecules for diverse applications in molecular biology, biotechnology, and medicine. The cornerstone of modern oligonucleotide synthesis is the solid-phase phosphoramidite method, pioneered in the 1980s by Marvin Caruthers and colleagues, building on Robert Bruce Merrifield's earlier invention of solid-phase , for which Merrifield received the 1984 . In this approach, synthesis occurs on an insoluble solid support, such as controlled pore glass beads, where the growing oligonucleotide chain is anchored at its 3' end. The process involves four repeating cycles per addition: detritylation to remove the 5'-dimethoxytrityl (DMT) , coupling of a protected nucleoside phosphoramidite monomer activated by , capping of unreacted chains with to prevent errors, and oxidation of the phosphite triester linkage to a stable phosphotriester using iodine. Following synthesis, the oligonucleotide is cleaved from the support and deprotected using ammonium hydroxide, yielding the final product with high purity if stepwise yields exceed 98.5%. Oligonucleotides synthesized via this method are indispensable tools in research, serving as primers for PCR amplification, probes for hybridization assays, and substrates for enzymatic studies. In therapeutics, they form the basis of antisense oligonucleotides (), small interfering RNAs (siRNAs), and splice-switching oligonucleotides (SSOs) that modulate to treat diseases ranging from rare genetic disorders to cardiovascular conditions. Notable FDA-approved examples include , an siRNA for , and several for and . Despite its efficiency, traditional phosphoramidite faces challenges in scalability, sustainability, and stereocontrol, particularly for phosphorothioate-modified backbones common in therapeutics, which generate diastereomeric mixtures and require large volumes of organic solvents like . Recent advances include liquid-phase alternatives with soluble supports, biocatalytic enzymatic for greener production, and chiral auxiliaries for stereoselective linkages, enabling larger-scale manufacturing up to multi-kilogram batches for clinical use. These innovations are driving the expansion of oligonucleotide-based drugs, with over a dozen approvals since 2016 and a robust pipeline targeting undruggable targets.

Introduction

Definition and scope

Oligonucleotide synthesis is the process of chemically or enzymatically assembling defined sequences of single-stranded nucleic acids, known as , from monomers. are short chains typically ranging from 10 to 100 in length and serve as fundamental building blocks in . The scope of oligonucleotide synthesis includes both deoxyribonucleic acid (DNA) and ribonucleic acid (RNA) sequences, as well as synthetic analogs incorporating modified nucleobases, sugars, or phosphate backbones to improve properties such as resistance to enzymatic degradation. Common modifications, like locked nucleic acids (LNA) or phosphorothioate linkages, enhance stability while maintaining hybridization specificity. Synthesis methods are broadly categorized into chemical and enzymatic approaches, with chemical methods often relying on iterative coupling of protected monomers and enzymatic methods using polymerases or transferases for sequence-directed assembly. A key feature of chemical synthesis is its 3' to 5' directionality, starting from the 3'-terminal nucleotide, whereas enzymatic synthesis proceeds in the natural 5' to 3' direction by adding nucleotides to the growing 3' end. These techniques enable the production of custom oligonucleotides for applications in research and therapeutics.

Applications in research and therapeutics

Synthetic oligonucleotides serve as essential tools in research, enabling precise manipulation and analysis of genetic material. In (PCR), short synthetic primers, typically 18-25 long, are designed to anneal to specific DNA sequences, facilitating amplification of target genes for downstream applications like cloning and sequencing. Hybridization probes, such as those used in (FISH) or technologies, allow detection of sequences through complementary base pairing, aiding in and mutation identification. Furthermore, synthetic oligonucleotides underpin gene synthesis by providing building blocks for assembling longer DNA strands via methods like , supporting efforts to genetic circuits and . In CRISPR-Cas systems, single-guide RNAs (sgRNAs) synthesized to match target DNA sequences direct Cas nucleases for precise , revolutionizing and disease modeling. These applications rely on the methods that allow incorporation of modifications for enhanced specificity and stability. In therapeutics, synthetic oligonucleotides have transformed treatment paradigms for genetic disorders and cancers through modalities like antisense oligonucleotides (ASOs), which bind mRNA to inhibit protein expression, and small interfering RNAs (siRNAs), which trigger RNA interference for gene silencing. Aptamers, structured RNA or DNA molecules selected for high-affinity binding to targets, offer alternatives to antibodies in diagnostics and targeted therapies. Components of mRNA vaccines, such as modified nucleosides incorporated during synthesis, improve translational efficiency and reduce immunogenicity, as demonstrated in COVID-19 vaccines. Notable FDA-approved examples include , the first approved in 1998 for in AIDS patients by inhibiting , and , an siRNA approved in 2018 for hereditary transthyretin-mediated , delivered via lipid nanoparticles to silence mutant TTR in the liver. These successes have expanded to 23 approved drugs as of November 2025, targeting neuromuscular, metabolic, and infectious diseases. The impact of oligonucleotide synthesis extends to and diagnostics, where custom-synthesized probes enable rapid for tailored therapies, such as companion diagnostics for cancer mutations. In diagnostics, facilitate point-of-care tests like those integrating for pathogen detection, enhancing precision in clinical decision-making. The oligonucleotide therapeutics market, driven by these innovations, reached approximately USD 6.2 billion in 2025, reflecting robust growth fueled by rising demand for gene-targeted treatments. Despite these advances, challenges in delivery and stability persist, as oligonucleotides are susceptible to nuclease degradation and poor cellular uptake, limiting efficacy outside the liver. These hurdles have spurred innovations in conjugation chemistries and formulations to improve and tissue targeting, thereby broadening therapeutic applicability.

Historical development

Early solution-phase methods

The pioneering efforts in oligonucleotide synthesis during the 1950s and 1960s relied on solution-phase approaches, with the phosphodiester method developed by H. Gobind Khorana marking a foundational advancement. This technique involved stepwise coupling of protected 3'-phosphates to the 5'-hydroxyl group of a growing chain, forming phosphodiester linkages central to natural nucleic acids. The activation was achieved using dicyclohexylcarbodiimide (DCC) to convert the phosphomonoester into a capable of phosphorylating the incoming nucleoside. Protecting groups were essential to prevent side reactions: the 5'-hydroxyl was typically shielded with a dimethoxytrityl (DMT) group, while exocyclic amines on bases were acylated using isobutyryl for and benzoyl for and . However, the method suffered from low coupling yields (often below 80%), prone to branching at phosphate sites due to over-activation, and required laborious chromatographic purification after each step, limiting practical synthesis to short oligomers of 10-15 . In the 1970s, the phosphotriester method emerged as a significant improvement, introduced by Robert L. Letsinger in 1969 and refined by groups including Keiichi Itakura and Saran A. Narang. This approach utilized 3'-O-phosphotriesters, where the was protected by two alkoxy groups—one linking to the previous nucleoside and a temporary group like 2-cyanoethyl for enhanced stability against compared to phosphodiesters. Coupling proceeded via activation of the trieseter with agents such as mesitylenesulfonyl chloride (MsCl) or arylsulfonyl azoles (e.g., mesitylenesulfonyl-3-nitro-1,2,4-triazole, MSNT), enabling the 5'-OH of the next protected nucleoside to displace the temporary and extend the chain. Itakura's group at the National Medical Center demonstrated efficient synthesis of deoxyoligonucleotides up to 20 units using this method, incorporating 2-cyanoethyl protection and selective deprotection under mild conditions. Narang's team at the National Research Council of Canada further optimized it by employing triazolides of arylsulfonic acids as activators, achieving coupling efficiencies around 95-97% and reducing side products. A general reaction for forming the phosphotriester intermediate is: \text{Nucleoside-OH} + (\text{RO})_2\text{P(O)Cl} \rightarrow \text{phosphotriester [intermediate](/page/Intermediate)} Despite these advances, repetitive and deprotection cycles remained time-consuming, and was constrained by issues in , often capping overall yields for longer sequences below 1%. Key milestones underscored the potential and challenges of these early methods. Khorana's group achieved the total of the 77-nucleotide structural gene for in 1970, assembling it from short via phosphodiester linkages and enzymatic joining, representing the first wholly synthetic gene and enabling functional studies in . This feat, involving over a decade of refinement, highlighted the method's utility for sequence-specific synthesis but also its labor intensity, as the project required synthesizing and purifying numerous fragments manually. By the late , phosphotriester approaches had facilitated the preparation of gene fragments for early experiments, yet persistent issues with yield accumulation and purification bottlenecks spurred the shift toward solid-phase techniques.

Transition to solid-phase approaches

The concept of solid-phase oligonucleotide synthesis was first introduced by Robert Letsinger in the 1960s, who demonstrated the attachment of nucleosides to insoluble beads, allowing sequential assembly without the need for extensive purification after each step. Although initial experiments yielded short oligomers, this approach laid the groundwork for immobilizing growing chains on a solid support, inspired by Merrifield's methodology. Practical adoption for oligonucleotide production accelerated in the 1970s, as improved protecting groups and coupling conditions enabled longer sequences and semi-automated processes. A key advancement in the late 1970s involved adapting the solution-phase phosphotriester method to solid supports, notably through the efforts of Marvin Caruthers and collaborators, who refined immobilization techniques to minimize intermediate isolations and enhance overall efficiency. This shift reduced the labor-intensive purification required in solution-phase syntheses, where each addition necessitated chromatographic separation of protected oligomers, by confining reactions to the resin-bound species and washing away byproducts. The phosphotriester approach on supports like controlled-pore glass or facilitated iterative assembly with step-wise yields of 95-98%, making it viable for sequences up to 20 . As a transitional method in the early 1980s, the H-phosphonate approach was developed by Brent Froehler, employing trivalent chemistry that permitted multiple couplings before a single oxidation step, bridging earlier phosphotriester limitations with emerging efficiencies. This method utilized stable H-phosphonate monomers, activated , to form P-H linkages that resisted premature oxidation, allowing chain extension in a convergent manner. A primary advantage was the elimination of intermediate isolations, as the entire synthesis cycle—phosphitylation of the support-bound 5'-hydroxyl, followed by of the incoming and repetitive chain extension—concluded with a global oxidation to yield phosphodiesters or phosphonates, streamlining the process and improving scalability for solid-phase formats.

Phosphoramidite method

Building blocks

The building blocks for the phosphoramidite method of oligonucleotide synthesis are primarily protected phosphoramidites, which serve as the monomers for sequential assembly of the oligonucleotide chain. These compounds feature a core with the 5'-hydroxyl group protected by a 4,4'-dimethoxytrityl (DMT) group to prevent unwanted reactions during synthesis, while the exocyclic amino groups on the bases are protected to avoid side reactions: N6-benzoyl for (dA), N4-acetyl or N4-benzoyl for deoxycytidine (dC), N2-isobutyryl for deoxyguanosine (dG), and no protection for (dT) or uridine (U). The key is at the 3'-position, where the hydroxyl is derivatized as a , specifically a 3'-O-(N,N-diisopropylamino)(2-cyanoethoxy)phosphinyl moiety. This P(III) structure consists of a atom bonded to the 3'-oxygen, the diisopropylamino group (which imparts reactivity), and the 2-cyanoethoxy group (which serves as a removable under basic conditions post-synthesis). The synthesis of these nucleoside phosphoramidites begins with commercially available or prepared protected s (5'-O-DMT, base-protected, 3'-OH free). Phosphitylation occurs by reacting the 3'-OH with chloro(N,N-diisopropylamino)phosphine in the presence of a base such as (DIPEA) in an anhydrous solvent like , yielding the intermediate 3'-O-bis(N,N-diisopropylamino)phosphino nucleoside. Subsequent cyanoethylation involves treatment with 3-hydroxypropionitrile (2-cyanoethanol) under similar conditions, displacing one diisopropylamino group to form the final 3'-O-(N,N-diisopropylamino)(2-cyanoethoxy)phosphinyl derivative. The product is purified by and isolated as a white foam, typically in 70-90% overall yield depending on the nucleoside. This two-step process ensures high purity and stereochemical integrity at the phosphorus center, though the phosphoramidites exist as a of diastereomers. Non-nucleoside phosphoramidites expand the versatility of the method by incorporating functional modifications into , such as spacers for chain extension, for purification, or fluorophores for detection. These are typically derived from non-nucleosidic scaffolds, like hexaethylene glycol for spacers or aminocaproic acid-linked , derivatized at one end with the standard moiety for incorporation during synthesis. For instance, phosphoramidites allow site-specific labeling without altering base pairing. While some are designed for pre-attachment to solid supports like controlled pore glass (CPG), their primary role here is as soluble modifiers. The reactivity of these building blocks stems from the P(III) center, which exhibits nucleophilicity toward activators like tetrazole, enabling efficient coupling in the synthesis cycle. They are stored as 0.1 M solutions in anhydrous acetonitrile under argon at -20°C to maintain stability for several months, as exposure to moisture or oxygen can lead to oxidation or hydrolysis.

Solid supports and linkers

Solid supports serve as the insoluble foundation for anchoring the initial nucleoside in solid-phase oligonucleotide synthesis, facilitating the stepwise assembly of the oligonucleotide chain while enabling efficient washing of reagents and byproducts. Common materials include controlled pore glass (CPG), which features rigid, non-swelling particles with pore sizes typically ranging from 500 to 1000 Å to accommodate the growing oligonucleotide chain within its matrix. CPG supports offer high mechanical stability and uniform pore distribution, making them suitable for automated synthesizers, with loading capacities generally between 30 and 100 μmol/g depending on pore size and functionalization. Alternative supports, such as highly cross-linked polystyrene beads, provide good moisture exclusion and swell in organic solvents like acetonitrile, enhancing reagent accessibility during synthesis. Tentagel, a composite of polystyrene grafted with polyethyleneglycol (PEG), combines the swelling properties of polystyrene with higher loading capacities (up to 200 μmol/g) and improved solvation for longer oligonucleotides. Linkers connect the 3'-end of the first to the solid , ensuring stable during synthesis and controlled release at the end. The succinyl linker, one of the most widely used, forms an bond that attaches the 3'-O-succinylated nucleoside to an amino-functionalized , such as aminopropyl-CPG, via amide coupling using activating agents like dicyclohexylcarbodiimide. This linkage is cleaved post-synthesis through base of the ester with concentrated ammonium hydroxide, typically at for 1-2 hours. Universal supports, which allow synthesis of any sequence from a single pre-functionalized material, often employ linkers like the PAC (phenoxyacetyl-protected) or similar variants that enable rapid cleavage and reduce preparation time by avoiding nucleoside-specific loading. These universal linkers typically rely on β-elimination mechanisms triggered by , yielding a universal 3'-hydroxyl or after cleavage. The attachment process begins with amino-derivatization of the , followed by of the succinylated to form the ester-amide bridge, which must withstand the acidic and nucleophilic conditions of the synthesis cycle while remaining labile to final deprotection. This setup anchors the , permitting repeated cycles of , oxidation, and deprotection with intermediate washes to remove unreacted species. Key advantages of these supports and linkers include high purity through facile purification and in settings, though limitations arise in large-scale where CPG fragility can lead to breakage under pressure, favoring more robust polymeric alternatives like Tentagel.

Synthesis cycle

The phosphoramidite method for oligonucleotide synthesis relies on a repetitive four-step cycle performed on a solid support, adding one per iteration in the 3' to 5' direction. This cycle—deblocking, coupling, capping, and oxidation—enables chain elongation, with each full cycle typically requiring 15-30 minutes, including washes. The overall yield for an n-mer is determined by the coupling raised to the power of the number of added, such that a % stepwise yields approximately 37% for a 50-mer. The approach, pioneered by Beaucage and Caruthers, uses protected building blocks that react under mild conditions. Deblocking initiates the cycle by removing the 5'-dimethoxytrityl (DMT) from the terminal , exposing the reactive 5'-hydroxyl. This is achieved with an acidic reagent, such as 3% () in (), under conditions at for about 50 seconds. The released orange trityl cation can be quantitatively monitored by its at 495 nm, confirming complete deprotection and allowing real-time assessment of progress. Coupling follows, where the free 5'-OH nucleophilically attacks the phosphorus(III) center of the incoming DMT-protected , activated by (typically 0.5 M in ). This nucleophilic substitution forms an internucleoside phosphite triester and proceeds with high efficiency, often exceeding 99%, in 1-5 minutes at . The key reaction is represented as: \text{5'-OH} + \text{(DMT-N-B)-P(III)-NR}_2 \xrightarrow{\text{[tetrazole](/page/Tetrazole)}} \text{5'-O-P(III)(OR)(O-B-N-DMT)} + \text{HNR}_2 where B denotes the , R the 3'-O-protecting group from the support, and NR_2 the dialkylamino . Unreacted 5'-OH groups are then capped to block further extension and minimize truncated products. This is accomplished using in a mixture with and (THF), often with N-methylimidazole as a catalyst, for approximately 30 seconds at . Acetylation ensures that any failed couplings do not propagate, thereby reducing sequence errors. The cycle concludes with oxidation, converting the labile phosphite triester to a stable triester. This is performed using 0.015 M iodine in a /pyridine/THF mixture (2:20:78) for about 45 seconds at . For introducing phosphorothioate linkages, sulfur-transfer reagents can replace iodine in this step, as detailed in specialized protocols. The cycle then repeats for the next until the desired length is reached.

Phosphorothioate modifications

Phosphorothioate (PS) modifications involve the replacement of a non-bridging oxygen atom in the phosphodiester backbone of oligonucleotides with a sulfur atom, enhancing resistance to nuclease degradation and improving pharmacokinetic properties for therapeutic applications. This modification is particularly crucial for antisense oligonucleotides (ASOs), where it extends serum half-life and facilitates tissue distribution without significantly impairing hybridization to target RNA or DNA. A prominent example is mipomersen, an FDA-approved ASO for familial hypercholesterolemia that incorporates PS linkages to achieve potent target inhibition while maintaining RNase H activation. In the phosphoramidite synthesis cycle, PS incorporation is achieved by modifying the oxidation step: after the coupling of the 3'-phosphoramidite monomer to the growing chain forms a triester , traditional iodine-based oxidation is replaced with a sulfurization using specialized . Common sulfurizing agents include 3H-1,2-benzodithiol-3-one 1,1-dioxide (Beaucage reagent), introduced for efficient sulfur transfer in , and phenylacetyl (PADS), which offers scalability and compatibility with large-scale production. The sulfurization occurs post-coupling, typically within 30-60 seconds in or mixtures, yielding the stable phosphorothioate triester directly and eliminating the need for subsequent oxidation. Standard building blocks are used without modification, as the sulfurization step operates on the , achieving coupling and sulfurization efficiencies of approximately 95-99%. The sulfurization reaction can be represented as: \text{R-O-P(III)(OR')}_2 + \text{S-reagent} \rightarrow \text{R-O-P(V)(S)(OR')}_2 + \text{byproducts} where the trivalent of the phosphite triester reacts with the source to form the pentavalent phosphorothioate triester. A key challenge with PS modifications is the introduction of a chiral center at each modified linkage, resulting in a diastereomeric of Rp and Sp configurations (approximately 1:1 ratio per site) for oligonucleotides with multiple PS bonds, leading to up to 2^n diastereomers for an n-PS oligo. These diastereomers exhibit subtle differences in physicochemical properties, such as binding affinity and resistance, but their similar migration behaviors complicate purification by standard techniques like HPLC or PAGE, often requiring specialized stereoselective or separation methods for therapeutic optimization.

Automation and scaling

Laboratory automation

Laboratory automation has revolutionized oligonucleotide synthesis by enabling efficient, reproducible production of custom sequences in research environments, primarily using the phosphoramidite method. The development of automated synthesizers began in the early 1980s with commercializing the first commercial instrument, the Model 380A, in 1983, which automated the stepwise addition of nucleotides. This was followed by the widely adopted Model 394 in 1991, which supported parallel synthesis across multiple columns and became a staple in labs for its reliability and ease of use. Automated synthesizers are broadly categorized into column-based and flow-through types. Column-based systems, exemplified by the ABI 394, pack solid supports like controlled-pore glass (CPG) into disposable columns where flow through the matrix to build the chain from 3' to 5'. In flow-through designs, such as modern benchtop instruments like the OligoNavigator, pumps drive directly across the support surface, reducing reagent volume and enhancing mixing efficiency for shorter cycle times. Both types support , with early models handling 1-4 channels and contemporary systems accommodating up to 96 channels for simultaneous synthesis of multiple sequences. The process is fully computer-controlled, involving precise delivery of for deblocking, , oxidation (or sulfurization), and capping steps, followed by automated washing and waste collection to minimize manual intervention. Each cycle, which adds one , typically requires 5-10 minutes, allowing a 20-30 mer to be completed in 2-5 hours. Key features include real-time monitoring via the dimethoxytrityl (DMT) assay, which quantifies the DMT cation released during deblocking to evaluate yields (ideally >98%), and intuitive software for input, parameter optimization, and run scheduling. These systems operate at scales of 0.2-10 μmol, yielding 1-50 mg of crude per , ideal for applications like primers and probes. Post-2020 advancements have focused on enhancing instruments with faster pumps, improved for reduced times, and better with purification modules, enabling higher throughput while maintaining small-scale for . These updates build on the foundational of the , streamlining what was once a labor-intensive manual process.

Large-scale production

Large-scale production of for therapeutic applications requires overcoming the limitations of by implementing robust industrial processes that yield kilogram to multi-kilogram quantities while adhering to (GMP) standards. This shift addresses the growing demand for (RNAi) drugs and antisense , where consistent high purity and scalability are essential for clinical and commercial supply. Key challenges in scale-up include managing and in expanded systems, as uneven flow in reactors can lead to channeling and incomplete reactions, while exothermic steps demand precise to prevent degradation. (CFD) modeling has been instrumental in optimizing these parameters, enabling simulations of reagent distribution and reaction kinetics to predict performance in larger setups. The has transitioned to oversized columns, often exceeding 100 liters in volume, to accommodate higher loadings of solid supports, though this introduces issues like pressure drops and nonlinear flow rates that reduce efficiency at scales beyond 10 kg per batch. To mitigate these hurdles, advanced techniques such as continuous flow synthesis have been adopted, utilizing modular reactors to enhance mass and , thereby improving efficiencies and reducing cycle times compared to traditional batch methods. reactors represent an emerging innovation, where oscillating flow suspends the solid support for uniform reagent access and easier scale-up, potentially supporting multi-kilogram outputs with minimal backpressure. GMP compliance is paramount, incorporating validated cleaning protocols and in-process controls to achieve oligonucleotide purity greater than 98%, ensuring removal of truncated species and chemical impurities suitable for parenteral administration. Recent advances between 2023 and 2025 have focused on enhancing phosphoramidite monomer purity to above 99%, which minimizes side reactions and impurity profiles during solid-phase synthesis, leading to higher overall yields and simplified downstream processing. These improvements, coupled with process optimizations, have driven manufacturing enhancements. Notable examples include Alnylam's expanded facilities in Norton, Massachusetts, which as of September 2025 enable multi-metric ton production of siRNA drug substances via scalable phosphoramidite chemistry. Yield optimization strategies, such as solvent recycling in closed-loop systems, have been implemented by contract manufacturers like Bachem, achieving over 30% reduction in acetonitrile consumption while maintaining high product quality.

Microarray and high-throughput synthesis

Microarray and high-throughput oligonucleotide synthesis involves methods that fabricate probes directly on substrates, enabling the parallel production of vast numbers of distinct sequences for genomic applications. Two primary approaches dominate: light-directed synthesis, pioneered by using on glass slides, and inkjet-based synthesis, employed by Agilent, which deposits reagents via non-contact piezoelectric printing. Both adapt the standard chemistry but incorporate modifications for spatial control, such as photolabile protecting groups in light-directed methods to enable site-specific deprotection. In the light-directed process, the synthesis surface is coated with photolabile-protected nucleoside phosphoramidites; masked (UV) light exposure selectively deprotects predefined regions, allowing coupling of the next to proceed only at those sites, with capping and oxidation steps completing each cycle. This iterative masking builds sequences across millions of features simultaneously, yielding arrays with up to 10^6 spots on a single chip. Inkjet synthesis, conversely, uses precise picoliter-scale deposition of deprotected phosphoramidites and other directly onto the , facilitating similar cycle-based extension without masks. Feature resolutions in these systems reach 5-10 μm, supporting high-density packing. These arrays are widely applied in and (SNP) detection, where probes hybridize to target nucleic acids for readout via scanning. Recent advances, particularly in 2024, have improved the efficiency and quality of photolithographic synthesis for microarrays, reducing cycle times and enhancing hybridization signals for probes up to around 30 mers.

Alternative methods

H-phosphonate synthesis

The H-phosphonate method for oligonucleotide synthesis employs H-phosphonate monoesters as building blocks, which feature a trivalent center with a characteristic P-H bond. These monomers, typically 5'-O-protected with dimethoxytrityl (DMT) groups and bearing standard base protections, are activated to facilitate chain elongation on a solid support. Unlike methods requiring immediate oxidation after each coupling, the H-phosphonate approach allows multiple condensations to build the entire chain as a series of H-phosphonate diesters before a single final oxidation step to form the phosphodiester backbone. This strategy was first reported in 1986, enabling the synthesis of DNA oligonucleotides up to 20 mers with yields comparable to early phosphoramidite protocols. The synthesis cycle begins with the acid-mediated detritylation of the 5'-DMT group on the growing chain, exposing a free 5'-hydroxyl. The nucleoside 3'-H-phosphonate monomer is then activated using pivaloyl chloride in the presence of a base like pyridine, forming a reactive mixed phosphonic-carboxylic anhydride intermediate. This intermediate undergoes nucleophilic attack by the chain's 5'-hydroxyl, yielding an H-phosphonate diester linkage: \text{Chain-5'-OH} + \text{DMT-5'-Nuc-3'-P(O)(H)-OH} \xrightarrow{\text{PivCl, pyridine}} \text{Chain-5'-O-P(O)(H)-O-3'-Nuc-5'-DMT} The reaction proceeds via a pyridinium adduct mechanism, exhibiting second-order kinetics dependent on the concentrations of the hydroxyl component and the H-phosphonate. Due to the relatively low reactivity of the H-phosphonate diesters, no capping step is required to block unreacted hydroxyls, simplifying the cycle. Multiple such couplings can be performed iteratively, typically achieving stepwise efficiencies of 95-98%. The full chain is then oxidized in a single step using iodine in water to convert all P-H linkages to stable phosphodiesters. For the preparation of phosphorothioate , the final oxidation is replaced by sulfurization using such as 3H-1,2-benzodithiol-3-one 1,1-dioxide or elemental , converting the H-phosphonate diesters to P-S bonds. This delayed modification step allows phosphorothioate incorporation across the chain and is particularly advantageous for analogs sensitive to stepwise sulfurization. The method supports solid-phase and has been scaled to produce multimilligram quantities of , such as a 14 μmol synthesis of a 15-mer with overall yields exceeding 50% after purification. Key advantages of the H-phosphonate method include the use of simpler, more stable reagents that avoid the moisture sensitivity of , making it suitable for synthesis where 2'-protecting groups must withstand acidic conditions during multiple couplings. It is also effective for longer (up to 50-60 mers) and high-throughput applications like microarrays, due to the stability of H-phosphonate linkages under basic conditions. However, limitations include slightly lower coupling efficiencies compared to modern methods, restricting its use for very long sequences (>100 mers) without optimization, and the need for a bulkier final oxidation step that can introduce side products if not controlled.

Enzymatic and biocatalytic approaches

Enzymatic and biocatalytic approaches to oligonucleotide synthesis represent an emerging paradigm that leverages biological catalysts to assemble nucleic acids, offering potential alternatives to traditional chemical methods by enabling template-independent polymerization in aqueous conditions. Central to these methods is the use of terminal deoxynucleotidyl transferase (TdT), a template-independent DNA polymerase that catalyzes the addition of deoxynucleoside triphosphates (dNTPs) to the 3'-hydroxyl (3'-OH) terminus of a growing DNA strand, allowing controlled, iterative extension without the need for protecting groups or harsh organic solvents. This process typically involves the sequential addition of reversibly blocked or activated dNTPs, followed by enzymatic or chemical unblocking to achieve single-base resolution, thereby mimicking natural polymerization while enabling de novo synthesis of custom sequences. A key advancement in enzymatic DNA synthesis (EDS) is the development of multiplex methods that permit parallel synthesis of multiple oligonucleotides with high fidelity. In 2023, researchers demonstrated a TdT-based approach using 3'-O-blocked dNTPs and a thermal unblocking step, enabling the simultaneous production of DNA strands up to 50 nucleotides long across 96 channels with over 98% accuracy per base, addressing scalability challenges in benchtop applications. Commercial systems like DNA Script's SYNTAX platform have operationalized this technology, automating the parallel enzymatic synthesis of up to 96 single-stranded DNA oligonucleotides, each up to 120 nucleotides in length, within hours using proprietary engineered TdT variants and microfluidic integration for reagent delivery and product elution. These systems highlight the process's efficiency, where TdT extends the primer's 3'-OH with a single activated dNTP per cycle, followed by deblocking to expose the new 3'-OH for the next iteration, yielding products suitable for downstream applications like gene assembly without extensive purification. Biocatalytic strategies extend beyond TdT to include engineered and ligases for more versatile synthesis, often incorporating chemoenzymatic hybrids that combine chemical nucleotide activation with enzymatic coupling to enhance specificity and length. For instance, modified DNA polymerases can incorporate non-natural or enable template-directed , while TdT variants have been evolved for higher processivity and reduced idling (uncontrolled addition). A 2025 review in ChemBioChem details how these hybrids, such as palladium-mediated deprotection paired with polymerase extension, facilitate the synthesis of modified up to 100 bases with incorporation efficiencies exceeding 95%, bridging chemical precision with biological speed. Recent progress from 2023 to 2025 has pushed the boundaries of length in enzymatic methods, with Ansa Biotechnologies achieving the synthesis of a 1,005-base using an optimized TdT system on a solid support, surpassing previous limits and enabling direct production of fragments for therapeutic and applications. These approaches offer environmental advantages, including water-based reactions that reduce by up to 90% compared to chemistry, and potential cost savings through catalytic reuse of enzymes, though challenges persist in error rates (typically 1-2% per base) and scaling beyond laboratory prototypes. Ongoing engineering of TdT and complementary enzymes aims to mitigate these limitations, positioning enzymatic as a greener, more accessible tool for on-demand production.

Post-synthesis processing

Cleavage and deprotection

Following the completion of the solid-phase synthesis cycle, the oligonucleotide chain attached to the solid support undergoes cleavage to release it into solution and deprotection to remove temporary protecting groups, yielding the crude, functional nucleic acid sequence. Cleavage is achieved by hydrolyzing the ester linkage of the succinyl or similar linker connecting the 3'-end of the oligonucleotide to the solid support, most commonly using concentrated ammonium hydroxide (28-30% NH₃ in water) or a mixture of ammonium hydroxide and methylamine (AMA, 1:1 ratio). For DNA oligonucleotides, treatment with concentrated ammonium hydroxide at 55°C for 8-16 hours effectively cleaves the linker while initiating deprotection; similar conditions apply to RNA, though milder variants like AMA at 65°C for 5-10 minutes are preferred to minimize degradation. This process typically provides post-cleavage yields around 90%, with losses primarily from incomplete hydrolysis or support dissolution. Deprotection occurs concurrently or sequentially, targeting two main classes of groups: nucleobase protections and phosphate protections. protecting groups, such as benzoyl on and or isobutyryl on , are removed via base-catalyzed during the or treatment, ensuring the heterocyclic rings are unmasked without altering the Watson-Crick pairing sites. Phosphate protections, specifically the 2-cyanoethyl groups on the phosphotriester intermediates, are eliminated through β-elimination under the same basic conditions, generating the native phosphodiester backbone and releasing as a byproduct. Side reactions, such as branch formation from unreacted 5'-hydroxyls, are largely prevented by the capping step during , maintaining in the linear chain. For RNA oligonucleotides, deprotection requires additional caution to avoid premature or incomplete removal of 2'-O-protecting groups, which could lead to chain cleavage or 2'/3'-phosphate migration. Harsh basic conditions are thus minimized; instead, after base and phosphate deprotection with ammonia or AMA, the 2'-silyl groups (e.g., tert-butyldimethylsilyl, TBDMS, or triisopropylsilyloxymethyl, TOM) are selectively removed using tetrabutylammonium fluoride (TBAF, 1 M), either in tetrahydrofuran at room temperature for up to 24 hours or in anhydrous DMSO at 65 °C for 2.5 hours, depending on the protecting group and protocol. This fluoride-mediated desilylation proceeds without significant degradation, preserving the 2'-OH functionality essential for RNA structure and activity.

Purification techniques

Following deprotection, the crude product contains a mixture of full-length sequences and impurities, necessitating purification to isolate the desired product for downstream applications. Purification techniques exploit differences in physicochemical properties, such as hydrophobicity, charge, and size, to separate full-length from truncated sequences and depurinated products, which arise from incomplete couplings and base loss during synthesis, respectively. Reverse-phase (RP-HPLC) is a widely used method for purification, particularly effective for initial separation of crude mixtures. It employs C18 stationary phases and elution gradients of in aqueous buffers with ion-pairing agents like , allowing separation based on hydrophobicity; the dimethoxytrityl (DMT)-on strategy retains the terminal DMT group on full-length sequences, enhancing their retention relative to shorter, DMT-off failure products. This approach removes truncated sequences by leveraging the incremental hydrophobicity added per , achieving high resolution for sequences up to 100 . Anion-exchange HPLC complements RP-HPLC by targeting the negatively charged backbone of , suitable for post-DMT removal purification. It uses strong anion-exchange resins and salt gradients (e.g., ) in or Tris buffers, with often added for denaturing conditions to unfold and prevent secondary structures that could impair separation. This method excels at resolving charge-based differences, effectively isolating full-length products from depurinated impurities and n-1 deletion sequences. For small-scale applications, (PAGE) provides base-pair resolution by size, enabling manual excision of full-length bands from denaturing gels, often yielding purities exceeding 90% for under 100 mers. , using membranes with molecular weight cutoffs around 3-10 , serves as a supplementary for concentration and desalting in low-volume preparations, though it is less selective for impurity removal compared to chromatographic methods. Preparative HPLC scales support production from milligrams to grams per run, with RP or anion-exchange formats enabling multi-gram yields for therapeutic development. Purity targets typically exceed 95% for research-grade oligonucleotides and 98% for good manufacturing practice (GMP) therapeutics, ensuring removal of process-related impurities like truncated and depurinated species to meet regulatory standards. Recent advances, including oligo-specific chromatography media optimized for modified backbones, have improved throughput and resolution for therapeutic-scale purification.

Characterization

Analytical methods

Analytical methods are essential for verifying the identity, purity, and quantity of following synthesis, ensuring they meet specifications for applications in research, diagnostics, and therapeutics. These techniques provide orthogonal assessments, combining structural confirmation with functional evaluation to detect impurities like truncated sequences (n-1 or n-x) or modifications. Common approaches include spectroscopic, chromatographic, and sequencing-based methods, often integrated for comprehensive . Mass spectrometry plays a central role in confirming oligonucleotide molecular weight and resolving structural modifications. time-of-flight (MALDI-TOF) mass spectrometry offers high sensitivity for direct analysis of crude synthesis products, producing singly charged ions that simplify spectra interpretation and enable precise mass determination up to several kilodaltons. (ESI) mass spectrometry, often coupled with liquid chromatography, generates multiple charge states for higher resolution, facilitating the identification of sequence variants and post-synthetic modifications such as or phosphorothioate linkages through (CID). For instance, ESI with ion-pair reversed-phase liquid chromatography (IP-RPLC) achieves over 90% sequence coverage, distinguishing isobaric impurities effectively. High-performance liquid chromatography (HPLC) combined with ultraviolet (UV) detection quantitates oligonucleotides by measuring absorbance at 260 nm, where nucleic acids exhibit maximum absorption due to their aromatic bases. Ion-pair reversed-phase HPLC (IP-RP-HPLC) separates full-length products from failure sequences based on hydrophobicity, with gradients of triethylammonium acetate (TEAA) and acetonitrile providing baseline resolution for oligos up to 60 nucleotides. Purity is assessed by peak integration, where sharp, symmetric peaks indicate high-quality synthesis, and quantitation relies on Beer-Lambert law-derived extinction coefficients. Capillary electrophoresis (CE), particularly capillary gel electrophoresis (CGE), complements HPLC by separating oligonucleotides by size and charge, offering superior resolution for length distribution analysis of shortmers and longmers in antisense oligonucleotides (ASOs) and small interfering RNAs (siRNAs). CGE uses polymer-filled capillaries under denaturing conditions to achieve efficiency up to 10^6 theoretical plates, enabling detection of impurities at low percentages. Sequencing methods verify the exact nucleotide composition, particularly for longer or complex oligonucleotides. Enzymatic Sanger sequencing serves as a gold standard for confirming sequences up to 1,000 bases, involving chain-termination with fluorescent dideoxynucleotides followed by capillary electrophoresis readout, achieving 99.9% accuracy for synthetic verification. For extended oligos or high-throughput needs, next-generation sequencing (NGS) platforms like Illumina provide deeper coverage, though they are less common for routine short-oligo checks due to overkill complexity. Nuclear magnetic resonance (NMR) spectroscopy, primarily in research settings, elucidates detailed three-dimensional structures and dynamics of oligonucleotides in solution using 1H, 13C, and 31P nuclei at natural abundance, without requiring isotopic labeling for modified sequences. It confirms base pairing, sugar puckering, and backbone conformations through resonance assignments and NOESY experiments. Functional assays assess the practical performance of oligonucleotides, such as hybridization stability. Melting temperature (Tm) analysis measures the thermal dissociation of duplexes formed with complementary strands, using UV absorbance at 260 nm to track during heating ramps, providing insight into binding strength influenced by sequence and modifications. For example, (LNA) modifications increase Tm by up to 5°C per substitution, enhancing duplex stability. Binding affinity assays, often via or fluorescence-based methods, quantify association constants (Ka) for probe-target interactions, verifying specificity in applications like aptamers or primers. These assays ensure synthesized oligos function as intended post-purification.

Quality control metrics

Quality control in oligonucleotide synthesis ensures the synthesized products meet standards for purity, yield, integrity, and reproducibility, particularly for applications in research and therapeutics. Purity metrics primarily focus on the percentage of full-length product (FLP), which represents the desired complete sequence, and the control of impurities such as n-1 deletions (shortmers missing one nucleotide). For research-grade oligonucleotides, a FLP purity exceeding 90% is typically required to minimize background noise in experiments like PCR or hybridization assays. In contrast, therapeutic oligonucleotides demand higher standards, with FLP purity often exceeding 95% to ensure safety and efficacy in clinical use. n-1 impurities are limited to less than 5% in total for both grades, as higher levels can compromise potency and increase off-target effects; regulatory thresholds emphasize identification if exceeding 1% and qualification if over 1.5% for impurity groups including n-1 sequences. Yield calculations distinguish between crude and purified products to assess . Crude yields represent the total output post-synthesis before purification, often 10-50% of theoretical based on , while purified yields are lower, typically 20-40% recovery after removal of truncated species. Quantification relies on extinction coefficients (ε) at 260 nm, calculated via nearest-neighbor models that sum base-specific contributions adjusted for sequence context, enabling accurate concentration determination via Beer's law: concentration = absorbance / (ε × path length). This approach ensures precise dosing for downstream applications without over- or underestimation. Integrity checks verify structural features critical to function, especially for modified oligonucleotides. In phosphorothioate (PS)-modified sequences used in therapeutics, chirality at phosphorus centers generates diastereomers (Rp and Sp configurations), and uncontrolled mixtures can affect binding affinity and nuclease resistance; quality control monitors diastereomeric ratios to ensure reproducible stereochemistry where stereoselective synthesis is applied. For injectable therapeutics, endotoxin levels must be below 5 EU/kg body weight (non-intrathecal) or 0.2 EU/kg (intrathecal), typically achieving <0.01 EU/mg in purified products to prevent pyrogenic responses. Regulatory standards from the FDA and guide therapeutic oligonucleotide quality, with 2024-2025 updates emphasizing comprehensive impurity profiling. The 's 2024 draft guideline, with public consultation closed on 31 January 2025, specifies purity assays via LC-MS, with an identification threshold of 1.0% and qualification threshold of 1.5% for unspecified impurities, based on toxicological data, and requires batch-specific justification for FLP and shortmer controls. Similarly, FDA's 2024 nonclinical safety guidance qualifies oligonucleotide-related impurities like n-1 through studies, aligning with ICH Q3A/B for process impurities, without fixed numerical thresholds but stressing control to safe levels informed by class effects. These frameworks prioritize risk-based specifications over rigid metrics. Batch variability assesses reproducibility across syntheses to maintain consistency. Metrics include (CV) for FLP purity (<5% across batches) and yield (±10% deviation), achieved through standardized solid-phase protocols and in-process controls like coupling efficiency monitoring. Low variability ensures therapeutic equivalence, with diastereomer distribution in PS oligonucleotides showing <2% batch-to-batch differences in stereoselective processes. Analytical data interpretation from prior characterization supports these metrics by confirming lot-to-lot uniformity.

References

  1. [1]
    Nucleic Acids Book - Chapter 5: Solid-phase oligonucleotide synthesis
    All oligonucleotides are made by phosphoramidite solid-phase synthesis. The steps in solid phase synthesis, and the preparation of reagents.
  2. [2]
    Modern approaches to therapeutic oligonucleotide manufacturing
    ### Overview of Oligonucleotide Synthesis from Science Article (DOI: 10.1126/science.adl4015)
  3. [3]
    Nobel Prize in Chemistry 1984
    - **Award**: Nobel Prize in Chemistry 1984 awarded to Robert Bruce Merrifield.
  4. [4]
    Large-Scale de novo Oligonucleotide Synthesis for Whole-Genome ...
    Jun 21, 2021 · In this review, we first summarize the status of large-scale oligonucleotide synthesis technologies including the basic methodology and large-scale synthesis ...
  5. [5]
    Laboratory Technologies | Computational Medicine Center at ...
    Primers are short oligonucleotides, 12 to 100 bases in length, with sequence ... nucleotides (dideoxyNTPs) that terminate DNA strand elongation. These ...
  6. [6]
    Modified Oligonucleotides: New Structures, New Properties, and ...
    Apr 26, 2021 · Modified oligo- and polynucleotides have become important tools in molecular biology, biochemistry, medicinal chemistry, diagnostic procedures ...
  7. [7]
    Chemical vs. Enzymatic DNA Synthesis Techniques | Ansa Bio
    Sep 8, 2025 · Compare chemical vs. enzymatic DNA synthesis. See how Ansa Bio's enzymatic method delivers faster, greener, and more precise DNA production.Step-By-Step Chemical... · Limitations Of Chemical... · Enzymatic Dna Synthesis...
  8. [8]
    CRISPR-based diagnostics | Nature Biomedical Engineering
    Jul 16, 2021 · The versatility, robustness and sensitivity of PCR have made this technology the most commonly used for the detection of DNA and RNA biomarkers.
  9. [9]
    Synthetic DNA applications in information technology - Nature
    Jan 17, 2022 · Its value lies in its characteristic attributes, namely Watson-Crick base pairing, array synthesis, sequencing, toehold displacement and ...Missing: guides | Show results with:guides
  10. [10]
    Multiplexed CRISPR technologies for gene editing and ... - Nature
    Mar 9, 2020 · Multiplexed CRISPR technologies, in which numerous gRNAs or Cas enzymes are expressed at once, have facilitated powerful biological ...
  11. [11]
    Incorporation of bridged nucleic acids into CRISPR RNAs improves ...
    Apr 13, 2018 · Off-target DNA cleavage is a paramount concern when applying CRISPR-Cas9 gene-editing technology to functional genetics and human ...
  12. [12]
    The RNA revolution in medicine: from gene regulation to clinical ...
    Aug 25, 2025 · We discuss how fundamental research in RNA biology has guided therapeutic development and review FDA-approved RNA-based drugs that are ...
  13. [13]
    Landscape of small nucleic acid therapeutics - Nature
    Mar 10, 2025 · In the early days of the chemical modification of nucleic acid drugs, primary agents were synthesized, and recent studies have shown that their ...
  14. [14]
    RNA-Based Therapeutics: From Antisense Oligonucleotides to ...
    Most oligonucleotide-based drugs are ASOs and siRNAs. However, there is potential that larger RNA species can be developed as therapeutics such as messenger ...
  15. [15]
    Antisense oligonucleotides: a novel Frontier in pharmacological ...
    Nov 17, 2023 · Antisense oligonucleotides (ASOs) are short single stranded synthetic RNA or DNA molecules, whereas double-stranded RNA nucleotide sequences are called small ...<|control11|><|separator|>
  16. [16]
    RNA-based therapeutics: an overview and prospectus - PMC
    Jul 23, 2022 · The first FDA-approved ASO medication (1998),. withdrawal in ... Three siRNA drugs (patisiran, givosiran, and lumasiran) have been approved ...
  17. [17]
    Clinical Pharmacokinetics of Approved RNA Therapeutics - PMC
    Jan 1, 2023 · This review focuses on FDA-approved drugs that have oligonucleotide structures and can access pharmacokinetic/pharmacodynamic data as published information.
  18. [18]
    RNA Therapeutics: A Healthcare Paradigm Shift - PMC
    ... vaccines, PCR, autoimmune disorders, therapeutic proteins, repurposing drugs ... Patisiran was the first siRNA-based medication to reach the market. It is ...<|separator|>
  19. [19]
    Oligonucleotide Synthesis Market worth $24.7 billion by 2030
    May 6, 2025 · The oligonucleotide synthesis industry is projected to grow from USD 10.5 billion in 2025 and to reach USD 24.7 billion by 2030, ...
  20. [20]
    Advances in oligonucleotide drug delivery - Nature
    including chemical modification, bioconjugation and ...
  21. [21]
    A Short History of Oligonucleotide Synthesis | TriLink BioTechnologies
    Gobind Khorana became interested in the synthesis of oligonucleotides. He introduced two concepts to the field that made possible the convenient synthesis of ...
  22. [22]
    [PDF] Nucleic acid synthesis in the study of the genetic code - Nobel Prize
    Proposed reaction sequence for the preparation of high-molecular-weight. RNA messengers and the subsequent in vitro synthesis of polypeptides of known amino.
  23. [23]
  24. [24]
    Deoxy oligonucleotide synthesis via the triester method
    K Itakura, N Katagiri, S A Narang, C P ... Phehylthio group as a protecting group of phosphates in oligonucleotide synthesis via phosphotriester approach.
  25. [25]
  26. [26]
  27. [27]
  28. [28]
  29. [29]
  30. [30]
    RAPID FUNCTIONALIZATION AND LOADING OF SOLID SUPPORTS
    Though, nucleoside-loaded CPG could be obtained using these procedures, there was wide variability in the loadings that ranged from 30 to 80 micromol/g starting ...
  31. [31]
    [PDF] Solid Phase Oligonucleotide Synthesis - Biotage
    Highly cross-linked polystyrene beads have the advantage of good moisture exclusion properties and they allow very efficient oligonucleotide synthesis, ...
  32. [32]
    Large scale synthesis of oligonucleotides via phosphoramidite ...
    Large scale synthesis of oligonucleotides ... oligonucleotides are synthesized on an aminopolyethyleneglycol derivatized polystyrene (TentaGel) support.
  33. [33]
    Attachment of Nucleosides and Other Linkers to Solid‐Phase ...
    Mar 1, 2013 · This unit provides protocols for all steps immediately involved in the preparation of conventional solid supports for routine oligonucleotide ...
  34. [34]
    Nucleoside Phosphoramidites Containing Cleavable Linkers - Pon
    Jan 1, 2006 · Cleavage mechanisms for linker phosphoramidites. (A) The linkage 6 formed by coupling 3 to LCAA-CPG is hydrolyzed at the succinyl ester ...
  35. [35]
    A Comparative Study of Commercially Available Universal Supports ...
    Standard oligonucleotide synthesis uses a solid support that contains the first nucleoside covalently bound to the support by a linker that is hydrolyzed during ...Universal Supports · Dna Synthesis On Universal... · Rna Synthesis On Universal...<|control11|><|separator|>
  36. [36]
    diacetic acid ('Q-linker') as a replacement for succinyl and oxalyl ...
    Nucleosides with this linker arm ('Q-linker') can be attached to solid phase supports and used for oligonucleotide synthesis in the same way as succinyl linker ...
  37. [37]
  38. [38]
    Deoxynucleoside phosphoramidites—A new class of key ...
    The development of a new class of nucleoside phosphites is described. These compounds are stable to normal laboratory conditions, are activated by mild acid ...
  39. [39]
    Review Reaction pathways and technologies of in vitro DNA synthesis
    Four-step cycle for oligonucleotide synthesis by the phosphoramidite method. Similar to phosphotriester and phosphite-triester methods, the phosphoramidite ...
  40. [40]
    Standard Protocol for Solid-Phase Oligonucleotide Synthesis using ...
    Coupling times vary depending on the phosphoramidite being used. Standard base coupling is typically 30s but modifiers tend to need much longer (typically 5- ...
  41. [41]
    Phosphorothioates, Essential Components of Therapeutic ...
    Phosphorothioates have found their usefulness in the general area of oligonucleotide therapeutic applications. Initially this modification was introduced ...
  42. [42]
    Clinical pharmacological properties of mipomersen (Kynamro ... - NIH
    The phosphorothioate modification enhances nuclease resistance as well and provides low affinity protein binding that facilitates distribution to peripheral ...
  43. [43]
    [PDF] 203568Orig1s000 - accessdata.fda.gov
    Jan 23, 2013 · Mipomersen is a second generation phosphorothioate oligonucleotide. The five bases on both the 5' and 3' ends are modified by attachment of ...
  44. [44]
    Phosphorothioates - an overview | ScienceDirect Topics
    Fortunately, the phophorothiate backbone can be efficiently, cheaply and easily achieved through modification of the oxidation step of oligonucleotide synthesis ...
  45. [45]
    Sulfur-Transfer Reagents for Synthesis of Oligonucleotide ...
    Jul 9, 2018 · Beaucage reagent (1),2) bis(phenylacetyl) disulfide (2) ... Sulfurization Efficiency with “Aged” Solutions of Phenylacetyl Disulfide (PADS).
  46. [46]
    Efficient Synthesis of Antisense Phosphorothioate Oligonucleotides
    Efficiency of phosphorothioate oligonucleotide syntheses could be improved by increasing the concentration of dichloroacetic acid in toluene to 10% from 3% ...
  47. [47]
    Unlocking P(V): Reagents for chiral phosphorothioate synthesis
    Aug 2, 2018 · The invention of a P(V)-based reagent platform for programmable, traceless, diastereoselective phosphorus-sulfur incorporation.
  48. [48]
    Chirality matters: stereo-defined phosphorothioate linkages at the ...
    Jul 15, 2021 · We report an efficient and simple method to introduce chiral PS linkages and demonstrate that Rp diastereomers at the 5′ end and Sp diastereomers at the 3′ end ...
  49. [49]
    DNA/RNA Synthesizer Model 394
    The first ABI synthesizer, Model 380A, shipped in 1983. This object, Model 394 introduced in 1991, was the second wave of Applied Biosystems's DNA/RNA ...
  50. [50]
    Synthetic DNA Synthesis and Assembly - PubMed Central - NIH
    The commonly used phosphoramidite synthesis chemistry consists of a four-step chain elongation cycle that adds one base per cycle onto a growing oligonucleotide ...
  51. [51]
    OligoNavigator - New Generation Benchtop Flowthrough ...
    OligoNavigator is a compact, pump-driven instrument for process development and small-scale synthesis of oligonucleotides in the scale 0.1-12 mmol. The ...Missing: post- 2020
  52. [52]
    Synthesizing your own oligonucleotides: FAQs - LGC
    Oct 3, 2023 · Oligo synthesis on all MerMade machines occurs in parallel, so it takes around the same amount of time to make 1 or 192 oligos. The timing is ...Missing: flow- | Show results with:flow-
  53. [53]
    [PDF] Introduction to the Synthesis and Purification of Oligonucleotides
    Modern synthesizers can routinely produce sequences of 150 or more nucleotides in usable amounts (≥10 µg). Several strategies can be employed to enhance the ...
  54. [54]
    Oligonucleotide Preparation Approach for Assembly of DNA Synthons
    Multicolumn or array synthesizers allow us to perform multiple syntheses simultaneously, which allows us to run synthetic processes in parallel and prepare ...Missing: flow- | Show results with:flow-<|separator|>
  55. [55]
    [PDF] Evaluating and Isolating Synthetic Oligonucleotides
    The DMT assay for determining stepwise yields is the spectrophotometric measure of the amount of DMT cation (the DMT group is the 5' hydroxyl protecting group) ...
  56. [56]
    [PDF] User Bulletin: ABI 392/4 Nucleic Acid Synthesizers
    Nov 23, 1999 · For syntheses at the 40-nmol scale: Use 0.05 M phosphoramidites and. Biotin Amidite, along with the 0.2-µmol scale cycle and an additional.
  57. [57]
    Glen Report 33-16: Product Offerings Update – K&A Synthesizer ...
    With 4 to 32 column positions, these relatively compact systems are capable of synthesis scales of up to 10 µmol. Many of our customers use these ...
  58. [58]
    How Oligo Synthesizers Are Changing Molecular Biology - Bio Basic
    Since there are a lot of improvements happening in technology, such synthesizers have become faster, and more precise, plus more affordable. What Is An Oligo ...Missing: post- 2020 pumps
  59. [59]
    How New Innovations Are Making Oligonucleotide Synthesis Faster ...
    Sep 3, 2025 · (2024) – Modern approaches to therapeutic oligonucleotide synthesis, Science. Reviews recent advances in both chemical and enzymatic methods ...Missing: post- pumps
  60. [60]
    Optimization and scale‐up of oligonucleotide synthesis in packed ...
    Jul 31, 2014 · A computational fluid dynamics (CFD) model for the analysis of oligonucleotide synthesis in packed bed reactors was developed and used to ...
  61. [61]
    Solid-phase oligonucleotide synthesis - Vapourtec
    However, these have low loadings and large quantities are needed for a scaled-up synthesis which leads to heat and mass transfer challenges. To increase ...Missing: fluid dynamics
  62. [62]
    Oligonucleotides manufacturing: scale-up & purification
    Jul 11, 2025 · Batch sizes are typically restricted to under 10 kg due to issues like nonlinear flow rates and reduced product purity as columns increase in ...
  63. [63]
    Oscillating fluidized bed oligonucleotide synthesizer - Google Patents
    A method and device for building an oligonucleotide on a solid phase resin within a filter reactor, wherein the method and device as used as a solid phase ...
  64. [64]
    Oligonucleotide therapeutics heading for tons of quantity in annual ...
    Sep 16, 2020 · Coupling reactions are greater than 98% efficient across manufacturing scales, and the technology provides control of reactions by precisely ...
  65. [65]
    Phosphoramidites for Oligonucleotide Synthesis - Data Insights Market
    Rating 4.8 (1,980) Aug 12, 2025 · Phosphoramidites for Oligonucleotide Synthesis Projected to Grow at XX CAGR: Insights and Forecasts 2025-2033.
  66. [66]
    The Enzymatic Approach as a Scalable, Sustainable Alternative
    May 15, 2025 · Cost and productivity gains​​ Phosphoramidite synthesis suffers from declining yields as sequence length increases, driving up the cost per gram ...<|separator|>
  67. [67]
    How Alnylam's Manufacturing Investments Bring RNAi Therapeutics ...
    Sep 30, 2025 · Alnylam's latest approach will enable multi-metric ton capacity of drug substance, which can be relied upon as our pipeline of investigational ...Missing: Lonza yield optimization recycling
  68. [68]
    Sustainable Approaches to Oligonucleotide Manufacturing - Bachem
    Sep 10, 2024 · By continuously recycling mixed fractions, we've achieved over 30% reduction in solvent use compared to traditional batch methods. This 24/7 ...Missing: Lonza Alnylam
  69. [69]
    Light-Directed, Spatially Addressable Parallel Chemical Synthesis
    High-density arrays formed by light-directed synthesis are potentially rich sources of chemical diversity for discovering new ligands that bind to biological ...
  70. [70]
    Light-directed synthesis of high-density oligonucleotide arrays using ...
    Here we demonstrate a new approach for synthesizing DNA probe arrays that combines standard solid-phase oligonucleotide synthesis with polymeric photoresist ...
  71. [71]
  72. [72]
    How DNA Microarrays are Built - Bitesize Bio
    Apr 4, 2025 · In situ-synthesized arrays are high-density oligonucleotide probe DNA microarrays, with Affymetrix GeneChip arrays being the most common. These ...
  73. [73]
    [2] The Agilent In Situ‐Synthesized Microarray Platform - ScienceDirect
    Agilent microarrays are made using in situ-synthesis via inkjet printing of nucleotide precursors, with a standard probe length of 60 nucleotides.<|control11|><|separator|>
  74. [74]
    Accelerated, high-quality photolithographic synthesis of RNA ...
    Jul 31, 2024 · In this study, we developed a series of RNA phosphoramidites with improved chemical and photochemical protection of the 2′- and 5′-OH functions.Results · Microarray Synthesis · Oligonucleotide Synthesis...<|control11|><|separator|>
  75. [75]
  76. [76]
  77. [77]
    Enzymatic Strategies for Next-Generation DNA Synthesis
    Sep 3, 2025 · Chemical DNA synthesis is the gold standard for producing synthetic oligonucleotides. Despite four decades of success, fragment length and ...
  78. [78]
    Multiplex enzymatic synthesis of DNA with single-base resolution
    Jul 7, 2023 · Enzymatic DNA synthesis (EDS) has the potential to be faster and give higher-purity DNA with none of the drawbacks (1, 2) of the conventional ...
  79. [79]
    Enzymatic DNA Synthesis | DNA Script
    The Syntax STX-200 System combined with the Hi-Fidelity Kits enables enzymatic synthesis of up to 120 nucleotides. “A major bottle neck in executing ...Missing: 120mer | Show results with:120mer
  80. [80]
    Syntax Product - DNA Script
    Introducing the SYNTAX nucleic acid production platform enabling same-day synthesis of custom oligos using enzymatic DNA synthesis technology.
  81. [81]
    Recent Advances in Biocatalytic and Chemoenzymatic Synthesis of ...
    Jan 24, 2025 · In this Section, we will discuss recent progress made in synthesis of base-modified oligonucleotides, XNAs, controlled enzymatic synthesis, and ...
  82. [82]
    Enzymatic DNA Synthesis: Going to Great Lengths | The Scientist
    Jun 13, 2024 · When chemical approaches for synthesizing longer DNA oligonucleotides plateaued, scientists turned to a one-of-a-kind enzyme.Missing: 1728 smooth- surface RSC
  83. [83]
    Glen Report 20.24: Deprotection - Volume 1 - Deprotect to Completion
    Deprotection using ammonium hydroxide is the most traditional method and dates back to the earliest days of oligonucleotide synthesis.
  84. [84]
    Methylamine deprotection provides increased yield of ...
    Use of methylamine or methylamine/ammonium hydroxide as a cleavage and deprotection reagent for the solid phase synthesis of Oligoribonucleotides has ...
  85. [85]
    Understanding Oligonucleotide Synthetic Yields
    Most oligonucleotides are ordered by the scale instead of the quantity, particularly the smaller scales of 15 μmole and lower, with no guarantee of final yield.
  86. [86]
    Nucleic Acids Book - Chapter 6: RNA oligonucleotide synthesis
    The TOM group is cleaved using 1 M TBAF in THF, without any degradation of the RNA oligonucleotide, and is stable to both basic and weakly acidic conditions.Solid-phase synthesis of RNA · Alternative RNA protecting...
  87. [87]
    [PDF] Deprotection Guide - Glen Research
    Deprotection using ammonium hydroxide is the most traditional method and dates back to the earliest days of oligonucleotide synthesis.
  88. [88]
    Purification and characterisation of oligonucleotides - ATDBio
    High-performance liquid chromatography (HPLC) is an efficient method for the analysis and purification of synthetic oligonucleotides.
  89. [89]
  90. [90]
    Overview and Recent Advances in the Purification and Isolation of ...
    Dec 16, 2024 · Therapeutic oligonucleotides are a novel chemical modality with unique characteristics that offer significant therapeutic potential.
  91. [91]
    Ion-pair reversed-phase high-performance liquid chromatography ...
    Reversed-phase (RP)-HPLC purification has become a popular technique for oligonucleotide purification with the introduction of the “trityl on” (“DMT on ...
  92. [92]
    [PDF] Purification of Oligonucleotides with Polymeric Reverse Phase ...
    Reverse phase chromatography with DMT-On is a preferred route since it is very effective at separating the DMT-On target oligonucleotides from the DMT-Off ...
  93. [93]
  94. [94]
    Oligonucleotides Isolation and Separation—A Review on Adsorbent ...
    Anion exchange using polymeric pellicular resin technology is a robust and stable methodology for N-1 failure sequence analysis for both unmodified and modified ...
  95. [95]
    [PDF] Preparative Anion Exchange Chromatography for the Purification of ...
    DNA oligonucleotides have a high degree of negative charge and therefore can be separated according to length with anion exchange chromatography. Using crude ...
  96. [96]
    Oligonucleotide Purification Methods for Your Research - US
    PAGE is the go-to method for establishing base-level resolution across short and/or long oligonucleotides and can routinely achieve greater than 90% purity of ...Choosing the right... · Reversed-phase cartridge... · High-performance liquid...Missing: ultrafiltration | Show results with:ultrafiltration
  97. [97]
  98. [98]
  99. [99]
    [PDF] Large scale purification – of oligonucleotides with ion exchange ...
    HPLC analysis confirmed (Fig. 1)89.1 % purity of the sample before the chromatographic purification. The target purity for the oligonucleotide was 95 %.
  100. [100]
    MALDI-TOF mass spectrometry: an emerging technology for ...
    MALDI-TOF MS has certain advantages over ESI-MS viz. (i) MALDI-TOF MS produces singly charged ions, thus interpretation of data is easy comparative to ESI-MS, ( ...
  101. [101]
    Advancements in the characterisation of oligonucleotides by high ...
    2022年4月10日 · This review covers the recent advancements in the analysis and characterisation of oligonucleotides in 2021 by high performance liquid chromatography mass ...
  102. [102]
    HPLC methods for purity evaluation of man-made single-stranded ...
    2019年1月31日 · Several HPLC interaction modes have been used successfully for the separation of an oligonucleotide mixture. While initially the mode of choice ...
  103. [103]
    Analysis of therapeutic nucleic acids by capillary electrophoresis
    Sep 20, 2022 · Capillary electrophoresis (CE) has proven to be an efficient separation technique and has been widely used for analyzing oligonucleotides and nucleic acids.
  104. [104]
    Guidelines for Sanger sequencing and molecular assay monitoring
    Feb 18, 2020 · This guidance is aimed at assisting laboratories in providing consistent, high-quality, and reliable sequence data when using Sanger-based ...Sanger Sequencing · Preparing Dna Template For... · Sequence Analysis<|separator|>
  105. [105]
    NMR Structure Determination for Oligonucleotides - Current Protocols
    Apr 2, 2018 · NMR spectroscopy is a versatile tool for determining the structure and dynamics of nucleic acids under solution conditions.Missing: Sanger NGS verification
  106. [106]
    Thermodynamic treatment of oligonucleotide duplex–simplex ...
    Generally the free energy of melting of an oligonucleotide duplex, D, into its constituent complementary single strands, S1 and S2, is obtained from the ...Missing: assays | Show results with:assays
  107. [107]
    Mechanism of the extremely high duplex-forming ability of ...
    Jul 18, 2023 · Mechanism of the extremely high duplex-forming ability of oligonucleotides ... binding affinity (ΔTm = +5°C/mod.) with the 2′,4′-BNA/LNA-modified ...Materials And Methods · Crystallization · Results And Discussion
  108. [108]
    Method of Oligonucleotide Purification - Bio-Synthesis
    Nov 25, 2013 · For example, some oligonucleotides require characterization by ion exchange chromatography to achieve extremely high purity greater than 98%.
  109. [109]
  110. [110]
    [PDF] Guideline on the Development and Manufacture of Oligonucleotides
    Jul 17, 2024 · This guideline addresses specific aspects regarding the manufacturing process, characterisation,. 56 specifications and analytical control ...Missing: fluidized bed reactors
  111. [111]
    Purification of crude DNA oligonucleotides by solid-phase extraction ...
    Typical yield of purified product is 60–95%. Final purity, measured by capillary gel electrophoresis, was found to be 90% or greater. Alternatively, highly pure ...Missing: standards | Show results with:standards
  112. [112]
    Oligo quantification—getting it right | IDT
    Apr 26, 2023 · The coefficient of extinction for any oligonucleotide sequence can be calculated using IDT's free OligoAnalyzer™ Tool which will also determine ...
  113. [113]
    Why and How to Control P-Chirality in Phosphorothioated ...
    Dec 6, 2024 · In this Perspective, we will highlight the importance of P-chirality in therapeutic oligonucleotides and its origin in the manufacturing process ...
  114. [114]
    What is Endotoxin-GenScript
    What Is the Maximum Acceptable Endotoxin Level · Drug (injectable, intrathecal) - 0.2 EU/kg body weight · Drug (injectable, non-intrathecal) - 5 EU/kg body weight ...
  115. [115]
    [PDF] Nonclinical Safety Assessment of Oligonucleotide-Based Therapeutics
    For the most recent version of a guidance, check the FDA guidance web page at https://www.fda.gov/regulatory-information/search-fda-guidance-documents. 3 ...<|separator|>
  116. [116]
    Emerging trends in the characterization of oligonucleotide ...
    ... batch-to-batch reproducibility. The number of PS linkages varies by the oligonucleotide classes, with antisense oligonucleotides (ASOs) containing a higher ...