Fact-checked by Grok 2 weeks ago

Nucleic acid structure

Nucleic acids are a class of large biomolecules essential to all known forms of , serving as the primary molecules for storing and transmitting genetic in cells and viruses. They consist of two main types: deoxyribonucleic acid (), which encodes the genetic instructions for protein and is primarily found in the of eukaryotic cells, and ribonucleic acid (), which plays diverse roles in protein , gene regulation, and other cellular processes. and are both polymers composed of repeating monomeric units called , linked together by phosphodiester bonds to form long chains. Each nucleotide monomer in nucleic acids comprises three key components: a nitrogenous base, a five-carbon sugar (ribose or deoxyribose), and one or more phosphate groups. In DNA, the sugar is deoxyribose, and the four nitrogenous bases are adenine (A), thymine (T), cytosine (C), and guanine (G); in RNA, the sugar is ribose, and thymine is replaced by uracil (U), with the other bases remaining the same. The sequence of these bases along the nucleic acid chain encodes genetic information, with specific base-pairing rules—A with T (or U in RNA), and C with G—enabling the complementary nature of strands. The most iconic structural feature of DNA is its double helix configuration, consisting of two antiparallel polynucleotide strands twisted around a common axis, stabilized by hydrogen bonds between complementary base pairs and hydrophobic interactions in the core. This right-handed helix, with a diameter of approximately 2 nm and 10 base pairs per helical turn (spaced 0.34 nm apart), was first proposed by and in 1953 based on X-ray diffraction data from and . In contrast, is typically single-stranded and flexible, allowing it to fold into intricate secondary structures such as hairpins, loops, and stems through intramolecular base pairing, which are crucial for its functional roles like in ribozymes or recognition in . These structural differences underpin the distinct biological functions of DNA as a stable genetic archive and as a versatile intermediary.

Basic components

Nucleobases

Nucleobases are the aromatic nitrogenous compounds that form the core informational components of nucleic acids, distinguishing DNA from RNA through specific variants. These molecules attach to sugar-phosphate backbones via glycosidic bonds and enable sequence-specific recognition through hydrogen bonding patterns. The primary nucleobases are classified into purines and pyrimidines based on their ring structures, with ionization properties governed by values that ensure neutrality at physiological . Purine nucleobases, and , possess a bicyclic comprising a six-membered ring fused to a five-membered ring, providing extended and rigidity. , chemically known as 6-aminopurine, features an amino group at position 6, while , or 2-amino-6-oxopurine, includes an amino group at position 2 and a group at position 6. Both exist predominantly in their amino-keto tautomeric forms under neutral conditions, with rare enol or imino tautomers occurring transiently and potentially influencing base pairing fidelity. The values for these purines—approximately 4.15 for (protonation at N1) and 9.2 for ( at N1-H)—position them as neutral species at 7, minimizing electrostatic repulsion in polymers. Pyrimidine nucleobases, cytosine (C), thymine (T) in DNA, and uracil (U) in RNA, are characterized by a single six-membered heterocyclic ring with nitrogens at positions 1 and 3. Cytosine is 4-amino-2-oxopyrimidine, bearing an amino group at position 4 and a keto group at position 2; thymine is 5-methyl-2,4-dioxopyrimidine, with keto groups at positions 2 and 4 and a methyl substituent at position 5; uracil is 2,4-dioxopyrimidine, identical to thymine but lacking the 5-methyl group. This methyl group in thymine enhances hydrophobic interactions and stability in DNA compared to uracil in RNA, contributing to distinct evolutionary roles in genetic storage versus expression. Their pKa values, around 4.5 for cytosine (protonation at N3), 9.7 for thymine, and 9.5 for uracil (deprotonation at N3-H), similarly favor neutral forms at physiological pH. The hydrogen bonding capabilities of these nucleobases dictate complementary pairing: forms two hydrogen bonds with or uracil via its N1 acceptor and N6-H donor pairing with the O4 and N3-H of T/U, respectively, while forms three hydrogen bonds with through its O6 and N1-H donors and N2-H donor interacting with cytosine's N3, O2, and N4-H. These patterns, illustrated in the base pair diagrams below, ensure specificity and stability in nucleic acid duplexes.
  • N6-H (A) ... O4 (T/U)
  • N1 (A) ... H-N3 (T/U)
Guanine-Cytosine Pair (3 H-bonds):
  • O6 (G) ... H-N4 (C)
  • N1-H (G) ... N3 (C)
  • N2-H (G) ... O2 (C)
Beyond the canonical set, rare modified nucleobases occur in specific contexts, such as hypoxanthine (C₅H₄N₄O; 1,7-dihydropurin-6-one, a deaminated adenine derivative) present as the nucleoside in the wobble position of certain tRNA anticodons, enabling flexible codon recognition during .

Sugars and phosphate backbone

The sugar-phosphate backbone forms the structural scaffold of nucleic acids, consisting of alternating deoxyribose (in DNA) or ribose (in RNA) sugars linked to phosphate groups via phosphodiester bonds. In DNA, the sugar is 2-deoxy-D-ribose, a pentose lacking a hydroxyl group at the 2' carbon position, while RNA incorporates D-ribose, which includes this 2'-OH group. Both sugars exist predominantly in the β-D-furanose (five-membered ring) conformation, with the anomeric carbon (C1') linked to the nucleobase via a β-glycosidic bond, ensuring a consistent orientation in the polymer chain. This furanose form provides rigidity to the backbone while allowing rotational flexibility around the C4'-C5' and C3'-O3' bonds. The absence of the 2'-OH in enhances DNA's by preventing intramolecular nucleophilic attacks that could disrupt the phosphodiester linkages, making DNA suitable for long-term genetic storage. In contrast, ribose's 2'-OH group increases RNA's susceptibility to but also imparts greater conformational flexibility, particularly in single-stranded regions, enabling diverse folding motifs essential for RNA's functional roles. This structural difference influences overall polymer dynamics: RNA tends toward A-form helices with a wider major groove due to the 2'-OH's steric and hydrogen-bonding effects, while DNA favors the more elongated B-form. Phosphodiester bonds are formed through condensation polymerization, where the 5'-phosphate group of one reacts with the 3'-OH group of another, eliminating a and creating a covalent linkage between the 5' carbon and 3' carbon across the . This unidirectional 5' to 3' defines the orientation of chains, with synthesis and replication processes proceeding exclusively in this direction. The resulting backbone is polyanionic, as each carries a negative charge at physiological ( ≈ 1-2), necessitating counterions for charge neutralization and structural integrity. Monovalent cations such as Na⁺ and K⁺ serve as primary counterions, coordinating directly with the negatively charged oxygens to screen electrostatic repulsion and stabilize the . studies reveal that Na⁺ interacts more strongly with groups due to its higher , often forming closer ion- contacts, whereas K⁺ prefers interactions with atoms in the grooves, influencing patterns and minor conformational adjustments. These ion-specific bindings are crucial for maintaining backbone and preventing aggregation in cellular environments. The 2'-OH group in RNA uniquely enables base-catalyzed hydrolysis of the via a mechanism, where the deprotonated 2'-O⁻ acts as a to attack the adjacent , forming a 2',3'-cyclic phosphate intermediate and cleaving the chain. This reaction proceeds efficiently under alkaline conditions ( > 7), with rate enhancements from general base catalysis, rendering RNA far less stable than DNA—phosphodiester bonds in DNA are approximately 100-200 times more resistant to such cleavage due to the missing 2'-OH. This inherent lability contributes to RNA's transient nature , contrasting with DNA's robustness.

Nucleotides and polymerization

Nucleotides consist of a nucleobase linked to a pentose sugar (ribose in RNA or deoxyribose in DNA) via a β-N-glycosidic bond, with one to three phosphate groups attached to the 5'-oxygen of the sugar, forming nucleoside monophosphates (NMPs), diphosphates (NDPs), or triphosphates (NTPs). The triphosphate forms are the primary substrates for nucleic acid synthesis: deoxyribonucleoside triphosphates (dNTPs) for DNA and ribonucleoside triphosphates (NTPs) for RNA, providing the energy needed for polymerization through cleavage of the high-energy phosphoanhydride bonds. In cells, NTP levels are maintained higher than NMP or NDP levels to support efficient synthesis, with kinases such as nucleoside diphosphate kinase catalyzing the transfer of phosphate from ATP to NDPs. Nucleic acid polymerization occurs enzymatically via DNA or RNA polymerases, which catalyze the template-directed addition of s to a growing chain. , first isolated by in 1956, incorporates dNTPs complementary to a DNA template, forming a between the 3'-hydroxyl of the primer terminus and the 5'-phosphate of the incoming dNTP, with concomitant release of (PPi) that drives the reaction forward. follows a similar mechanism but initiates without a primer, using NTPs to synthesize RNA complementary to a DNA template starting from a promoter sequence, also releasing PPi; this process was elucidated through studies of bacterial enzymes like E. coli . Both enzymes require a template strand to ensure base-pairing specificity, with the incoming selected via hydrogen bonding to the template base in the . Polymerization proceeds exclusively in the 5' to 3' direction, where new are added to the 3'-hydroxyl end of the chain, resulting in a linear with 5' and 3' hydroxyl termini. This directionality arises from the chemical mechanism of nucleophilic attack by the 3'-OH on the α- of the incoming NTP or dNTP, preventing 3' to 5' . The release of is often hydrolyzed by pyrophosphatases to shift the toward polymer elongation. DNA and RNA synthesis differ in substrates and fidelity: DNA polymerases use dNTPs lacking the 2'-hydroxyl group, enabling a more stable double , while RNA polymerases incorporate rNTPs with the 2'-OH, which introduces greater flexibility but higher reactivity. DNA polymerases possess 3' to 5' proofreading activity, achieving error rates of approximately 10^{-7} to 10^{-9} per , whereas RNA polymerases lack robust , resulting in higher error rates of about 10^{-4} to 10^{-5}, suitable for transient RNA molecules.

Primary structure

Nucleotide sequence and composition

The primary structure of nucleic acids is defined as the precise linear sequence of nucleotides, determined by the specific order of their nitrogenous bases—adenine (A), cytosine (C), guanine (G), and thymine (T) in deoxyribonucleic acid (DNA), or uracil (U) replacing thymine in ribonucleic acid (RNA)—connected via phosphodiester bonds in the 5' to 3' direction. This sequence encodes genetic information and is conventionally denoted as a string of single-letter symbols, such as 5'-ATGC-3' for a short DNA segment. The nucleotide composition, particularly the GC content (the percentage of guanine and cytosine bases), significantly affects the stability and thermal properties of the nucleic acid. In double-stranded DNA, higher GC content correlates with increased melting temperature (Tm), the point at which half the double helix dissociates into single strands, due to the stronger hydrogen bonding between G-C pairs compared to A-T pairs. For oligonucleotides shorter than 20 bases under standard PCR conditions (e.g., 50 mM monovalent salt), an approximate Tm is given by the Wallace rule: T_m = 2 \times (A + T) + 4 \times (G + C) °C. Chargaff's rules describe the equimolar base ratios observed in most double-stranded DNA molecules: the quantity of adenine equals thymine (A = T), and guanine equals cytosine (G = C), arising from complementary base pairing along the two strands. These parity relationships, established through biochemical analyses of DNA from various organisms, do not apply universally; exceptions occur in single-stranded DNA or certain viral genomes where base pairing is absent or incomplete. Sequence motifs represent recurring patterns within the primary structure, such as simple tandem repeats. A prominent example is the poly-A tail in eukaryotic (mRNA), a homopolymeric stretch of 50–250 residues added post-transcriptionally at the 3' end.

Chemical modifications and stability

Chemical modifications to the primary structure of nucleic acids involve the addition of functional groups to nucleobases or the sugar-phosphate backbone, altering their chemical properties without changing the underlying sequence. In DNA, one of the most prevalent modifications is (5mC), which occurs primarily at CpG dinucleotides and plays a central role in epigenetic regulation by influencing through and transcriptional repression. This modification is catalyzed by DNA methyltransferases and is essential for processes such as genomic imprinting and X-chromosome inactivation. Another significant DNA modification is N6-methyladenine (m6A), which is widespread in bacterial genomes where it contributes to restriction-modification systems that protect against foreign DNA invasion. In bacteria like oryzae, m6A is installed by methyltransferases such as and helps regulate replication and repair pathways. In RNA, over 170 distinct chemical modifications have been identified, particularly in (rRNA) and (tRNA), where they fine-tune structure and function. Key examples include (Ψ), formed by isomerization of , which enhances base stacking and hydrogen bonding stability; N6-methyladenosine (m6A), the most abundant internal modification in eukaryotic mRNA that affects splicing, export, and translation; and 2'-O-methylation (Nm), which protects against degradation and modulates assembly. These modifications are enzymatically installed by writer proteins, such as pseudouridine synthases (PUS enzymes) that catalyze the reversible C-C glycosidic bond formation in Ψ without requiring cofactors. For instance, families like TruA and TruB in and Pus1-Pus10 in eukaryotes target specific sites in tRNA and rRNA. These modifications significantly impact nucleic acid stability by conferring resistance to enzymatic degradation and modulating helical properties. In therapeutic applications, such as (siRNA), incorporation of 2'-fluoro substitutions at the 2' position of the sugar enhances resistance, allowing prolonged activity while maintaining efficacy. Similarly, 5mC in DNA reduces backbone flexibility, increasing helix rigidity and protecting against hydrolytic cleavage. In RNA, Ψ and Nm stabilize secondary structures by improving thermodynamic stability and shielding against endonucleases like RNase A. Detection of these modifications relies on specialized techniques that preserve and identify the altered bases. For 5mC in DNA, is a cornerstone method that converts unmethylated cytosines to uracils via sulfonation and , while 5mC remains resistant, enabling differentiation through subsequent amplification and sequencing. This approach provides genome-wide mapping but requires careful optimization to minimize DNA fragmentation. Base composition, particularly CpG density, influences the prevalence of modifiable sites like those for 5mC.

Secondary structure

Base pairing rules and hydrogen bonding

In nucleic acids, base pairing refers to the specific association between nucleobases that stabilizes secondary structures through ing. In DNA, the canonical Watson-Crick base pairing rules dictate that (A) pairs with (T), and (G) pairs with (C). These pairings occur between a on one strand and a on the complementary strand, ensuring geometric uniformity in the double helix. The specificity arises from complementary donor and acceptor sites on the bases, which form precise interactions: the A-T pair involves two hydrogen bonds, while the G-C pair forms three, contributing to greater stability in G-C rich regions. The hydrogen bonds in these pairs are typically N-H···O or N-H···N types, involving the Watson-Crick faces of the bases. For A-T, one bond forms between the N1 of (donor) and N3 of (acceptor), and the second between the amino group at C6 of (donor) and the carbonyl at C4 of (acceptor). In G-C pairing, the three bonds are: N1 of to N3 of , the amino group at C2 of to the carbonyl at C2 of , and the amino group at C4 of (donor) to the carbonyl at C6 of (acceptor). These interactions not only dictate pairing fidelity but also influence melting temperatures, with each additional G-C bond increasing duplex stability by approximately 1-2 kcal/mol compared to A-T. In RNA, the base pairing rules are analogous but substitute uracil (U) for , forming A-U pairs with two s (N1 of to N3 of uracil, and amino at C6 of to carbonyl at C4 of uracil) and retaining G-C pairs with three. RNA often adopts single-stranded conformations with intramolecular base pairing to form stems in hairpin loops or other motifs, where patterns remain similar but allow for greater flexibility. The G-C pair's stronger bonding (due to the third ) promotes more stable RNA duplexes, as evidenced by higher thermal denaturation temperatures in GC-rich sequences. Beyond strict Watson-Crick pairing, non-canonical interactions like wobble base pairs occur, particularly in . Proposed by , the wobble hypothesis describes relaxed base at the third position of codons during , allowing a single tRNA to recognize multiple synonymous codons. For instance, in the anticodon can pair with either or uracil in the mRNA via two hydrogen bonds, shifting the geometry to accommodate the "wobble" without disrupting overall specificity. G-U wobble pairs, common in structures, feature two hydrogen bonds (N1 of G to O2 of U, and O6 of G to N3 of U) and introduce functional diversity, such as in where they influence decoding accuracy and structural dynamics. These wobble interactions are weaker than canonical pairs but essential for the degeneracy of the , reducing the required number of tRNAs from 61 to about 40.
Base PairMoleculeHydrogen BondsKey Donors/Acceptors
A-TDNA2A(N1)-T(N3); A(N6)-T(O4)
G-CDNA/RNA3G(N1)-C(N3); G(N2)-C(O2); G(O6)-C(N4)
A-URNA2A(N1)-U(N3); A(N6)-U(O4)
G-U (wobble)RNA2G(N1)-U(O2); G(O6)-U(N3)
This table summarizes the primary base pairs, highlighting the hydrogen bonding patterns that underpin nucleic acid structure.

DNA double helix configurations

The DNA double helix, formed through complementary pairing between adenine-thymine and guanine-cytosine, manifests in several distinct configurations that influence its overall geometry and biological function. These variants arise primarily from differences in backbone conformation, stacking, and hydration levels, with the B-form representing the predominant structure under physiological conditions. B-DNA is a right-handed characterized by a smooth, elongated structure with approximately 10.5 pairs per turn, a helical pitch of 3.4 nm, a rise of 0.34 nm per pair, and a twist angle of 36° between adjacent pairs. This configuration features distinct major and minor grooves, which facilitate interactions with proteins for processes such as replication and transcription. The structure was first proposed by and Crick based on data, with refined parameters derived from studies. A-DNA, also right-handed but shorter and wider than B-DNA, adopts a more compact form with 11 s per turn, a of about 2.8 , a rise of 0.23 per , and a twist of approximately 33°. In this conformation, the base pairs are tilted relative to the helix axis, resulting in a deep, narrow minor groove and a shallow major groove. A-DNA is favored under low-humidity conditions, such as in dehydrated fibers, and is commonly observed in DNA-RNA hybrids. Z-DNA represents a left-handed with a zig-zag backbone, accommodating 12 s per turn, a of 4.5 nm, a rise of 0.37 nm per , and a twist of -30°. Unlike the right-handed forms, Z-DNA has a single deep groove and no distinct major/minor distinction, with glycosidic conformations for purines and for pyrimidines. This form is stabilized in sequences rich in alternating purine-pyrimidine tracts, particularly GC repeats, and was first identified through crystallographic analysis of synthetic . The prevalence of these helical forms is modulated by environmental factors, including , , and sequence composition. B-DNA predominates in aqueous, physiological environments with moderate salt concentrations (e.g., ~150 mM NaCl), while A-DNA emerges at relative humidities below 75% or in the presence of alcohols that reduce . Z-DNA formation is promoted by high salt concentrations (e.g., >2 M NaCl) or multivalent cations like Mg²⁺, which screen repulsions, and is further enhanced in negatively supercoiled contexts or by specific protein binding, though the latter influences are secondary to ionic effects. Sequence motifs, such as AT-rich regions favoring B-DNA stability through optimal stacking, and GC-rich segments predisposing to Z-DNA via favorable syn-anti alternations, also play a key role. variations can induce transitions, with acidic conditions occasionally stabilizing A-like forms by protonating bases and altering hydrogen bonding.
Helix TypeHandednessBase Pairs per TurnPitch (nm)Rise per Base Pair (nm)Twist Angle (°)Key Features
B-DNARight10.53.40.3436Major/minor grooves; physiological form
A-DNARight112.80.2333Tilted bases; low humidity
Z-DNALeft124.50.37-30Zig-zag backbone; high salt/GC-rich

RNA folding motifs

RNA folding motifs are local secondary structural elements that arise from base pairing within single-stranded RNA molecules, enabling diverse functions such as , , and molecular . Unlike the continuous double helix of DNA, these motifs feature discontinuous helical regions interrupted by unpaired , forming compact architectures stabilized by hydrogen bonding and stacking interactions. The most fundamental motif is the stem-loop, consisting of a double-stranded helical stem formed by complementary base pairing and an unpaired loop of 4-7 at the apex. Stem-loops can be further diversified by bulges and internal loops, where unpaired protrude from one or both strands, respectively, disrupting the continuity of the and introducing flexibility or sites. For instance, bulge loops with a single unpaired on one strand facilitate sharp turns in the backbone, while internal loops with unpaired residues on both strands allow for asymmetric expansions that accommodate tertiary contacts. Hairpins represent a specific class of stem-loops where the loop size and sequence confer exceptional , particularly tetraloops (four-nucleotide loops) with sequences like GNRA, which exhibit enhanced thermodynamic due to non-canonical base interactions and stacking. These tetraloops are among the most stable loop configurations, with contributions up to 4-6 kcal/ more favorable than larger loops, as determined from optical studies. Magnesium ions (Mg²⁺) further stabilize these motifs by bridging negatively charged groups in loops and stems, reducing electrostatic repulsion and promoting compact folding, especially in bulged or internal regions. Beyond simple stem-loops, pseudoknots form when a single-stranded region base-pairs with a complementary outside an existing stem, creating interleaved helices that cross over like a and often enhance mechanical rigidity or signaling. Kissing loops occur when the apical loops of two separate hairpins interact via complementary base pairing, forming transient or stable intermolecular contacts that mediate dimerization or regulatory switching. In transfer RNA (tRNA), the cloverleaf secondary structure exemplifies the integration of multiple stem-loops, including the acceptor stem, D-loop, anticodon arm, and T-loop, which collectively form four helical arms connected by loops for amino acid attachment and codon recognition. Riboswitches, regulatory RNA elements in bacterial mRNAs, frequently incorporate pseudoknots and kissing loops alongside stems to sense metabolites like thiamine or guanine, undergoing conformational changes that control gene expression. Computational tools such as mfold and the ViennaRNA package enable prediction and identification of these s by minimizing using dynamic programming algorithms that account for base-pairing rules, penalties, and stacking energies. Mfold, developed by Zuker, computes optimal and suboptimal foldings for sequences up to several hundred , while ViennaRNA extends this with advanced features like prediction and covariance models for detection in alignments.

Tertiary structure

DNA supercoiling and topology

DNA supercoiling represents a key aspect of the tertiary structure of closed circular DNA molecules, where the double helix, serving as the substrate, undergoes additional coiling beyond its intrinsic helical twist to achieve compaction and facilitate biological processes. In covalently closed circular DNA, such as bacterial plasmids or viral genomes, the topology is invariant unless broken and resealed, leading to superhelical tension that influences DNA accessibility and function. The topological state of supercoiled DNA is quantified by the (Lk), defined as the sum of the (Tw), which measures the helical turns of the two strands around each other, and the writhe (Wr), which captures the coiling of the helical axis in space:
\mathrm{Lk = Tw + Wr}
This relationship, established through mathematical analysis of ribbon topology, holds for any closed DNA duplex. The relaxed linking number (Lk₀) corresponds to the state without supercoiling, typically about 10.5 base pairs per turn in B-form DNA. Supercoiling arises when Lk deviates from Lk₀, quantified by ΔLk = Lk - Lk₀; negative ΔLk indicates underwinding (negative supercoiling), while positive ΔLk indicates overwinding (positive supercoiling). Negative supercoiling predominates , promoting DNA unwinding for processes like replication and transcription, whereas positive supercoiling can accumulate ahead of progressing polymerases.
To manage superhelical tension, cells employ DNA topoisomerases, enzymes that transiently break and rejoin DNA strands to alter . Type I topoisomerases, such as the enzyme discovered in 1971, relax supercoils by nicking one strand, changing in steps of ±1 without requiring ATP; they preferentially relieve negative supercoils. Type II topoisomerases, including , act on both strands, altering in steps of ±2 and often requiring ATP; while most type II enzymes relax supercoils bidirectionally, gyrase uniquely introduces negative supercoils using ATP hydrolysis, counteracting the positive supercoils generated during transcription. In , gyrase maintains an overall negative superhelical density (σ ≈ -0.06) essential for chromosomal compaction within the confined space. Supercoiled DNA adopts distinct three-dimensional configurations to partition the writhe component. Plectonemic supercoils form right-handed interwound structures where the DNA axis coils around itself, typical in unconstrained bacterial DNA and allowing dynamic partitioning of twist and writhe. In contrast, toroidal (or solenoidal) supercoils involve the DNA wrapping left-handedly around a core, as seen in eukaryotic nucleosomes where approximately 147 base pairs wrap 1.65–1.7 turns around the histone octamer, contributing about -1 supercoil per nucleosome. This wrapping constrains negative supercoils, reducing free writhe and aiding chromatin compaction. In bacteria, negative supercoiling driven by gyrase compacts the genome by favoring plectonemic structures and branched domains, while also linking to replication by relieving torsional stress at forks and to transcription by enhancing promoter opening and RNA polymerase progression.

RNA three-dimensional domains

RNA molecules fold into compact three-dimensional structures that integrate secondary structural motifs, such as and loops, to form functional tertiary domains essential for biological roles like and molecular recognition. These tertiary folds are stabilized by long-range interactions, including base triples, coaxial stacking of , and coordination with metal ions like Mg²⁺, which neutralize phosphate repulsions and position catalytic groups. Unlike the more rigid DNA double helix, RNA's single-stranded nature allows diverse architectures, often involving non-canonical base pairs and pseudoknots that pack secondary elements into globular shapes. A quintessential example of RNA tertiary structure is transfer RNA (tRNA), which adopts a conserved L-shaped fold approximately 7 nm long and 2 nm wide. This architecture arises from the coaxial stacking of two orthogonal helical domains: the acceptor stem and T-arm form one arm of the L, while the D-arm and anticodon arm stack to form the other, with the anticodon loop at the distal end and the 3' acceptor site at the opposite tip. The fold is further stabilized by tertiary interactions, such as hydrogen bonds between the and T-loop, and Mg²⁺ ions that bridge phosphates in the core, enabling tRNA's role in translation by positioning the anticodon for mRNA decoding and the acceptor for aminoacylation. The atomic-resolution of yeast tRNA^Phe, determined at 1.93 Å resolution, revealed these features, highlighting the universal tertiary organization across tRNAs despite sequence variability. Ribozymes exemplify how RNA tertiary domains create active sites for catalysis, mimicking protein enzymes. The hammerhead , a small self-cleaving found in viroids and RNAs, folds into a Y-shaped with three helices flanking a conserved core of 13-15 that positions the scissile for inline attack by 2'-OH. Crystal structures of the hammerhead in pre-cleavage and post-cleavage states show a compact geometry where a sheared G-A and a turn organize catalytic residues, enabling site-specific with rate enhancements up to 10^6-fold over uncatalyzed cleavage. Similarly, group I introns, larger s (250-500 ) in organelles and , fold into a barrel-like tertiary domain with two stacked helical domains and peripheral elements that form a guanosine-binding site adjacent to the 5' splice site. This architecture, resembling the spliceosome's catalytic core, facilitates two steps for self-splicing, with the geometry involving a U-G wobble pair and coordinated Mg²⁺ ions to activate the ; crystallographic studies of the Azoarcus group I intron at 3.1 confirmed these interactions, underscoring evolutionary conservation. Tertiary packing in RNA often relies on modular interactions between secondary motifs, such as tetraloop-receptor pairings and kissing loops, which enhance stability and specificity. GNRA-type tetraloops (e.g., ) bind to internal receptor loops via non-canonical base pairs and hydrogen bonds, as seen in the P4-P6 domain of the group I , where the tetraloop docks into a G-C rich receptor to rigidify the fold and boost folding rates by orders of magnitude. Kissing loops, formed by Watson-Crick pairing between complementary hairpin loops, mediate intermolecular or intramolecular contacts; for instance, in ribosomal RNAs, these loops pack helices coaxially, contributing to domain assembly, while in viral RNAs like HIV-1 DIS, they initiate dimerization with dissociation constants in the nanomolar range. These interactions, first structurally characterized in the at 2.8 Å resolution, exemplify how RNA tertiary domains achieve architectural modularity. Predicting RNA tertiary structures remains challenging due to the landscape of multiple folding minima and electrostatic complexities, but computational tools have advanced significantly. Early methods like Rosetta's FARFAR protocol and (MD) simulations achieved median RMSDs of 4-6 Å for blind predictions of systems up to 100 . More recent AI-based approaches, such as AlphaFold3 (2024), have improved accuracy for RNA structures, often achieving RMSDs below 3 Å for diverse domains by integrating multimodal data and , as demonstrated in benchmarks up to 2025. These advances facilitate modeling of complex RNA architectures, though challenges in long-range interactions persist.

Quaternary structure

Nucleic acid multimers and assemblies

Nucleic acid multimers and assemblies refer to quaternary structures where multiple strands of DNA or RNA interact to form complex, non-covalent architectures beyond simple duplexes, often playing roles in genetic recombination, gene regulation, and viral packaging. These structures exhibit higher-order symmetry and branching, stabilized by base stacking, hydrogen bonding, and metal ion coordination, without involvement of proteins. Examples include four-stranded DNA junctions and stacked guanine tetrads, which can adopt dynamic conformations influenced by sequence and environmental factors. In DNA, Holliday junctions represent a canonical four-stranded multimer formed during , consisting of two homologous duplexes connected at a crossover point to create a branched, X-shaped topology. This structure arises from reciprocal strand exchange between aligned DNA molecules, enabling branch migration and resolution to facilitate genetic exchange or repair. Crystal structures reveal that Holliday junctions adopt a stacked conformation with two continuous helices, where the junction angle is approximately 60 degrees, promoting stability through coaxial stacking of base pairs. Beyond recombination, similar four-way junctions occur in branched DNA intermediates, such as those resolved by endonucleases during DNA processing. G-quadruplexes exemplify another DNA multimer, formed by stacking of multiple G-tetrads—planar quartets of four guanine bases linked via Hoogsteen hydrogen bonding—in guanine-rich sequences. Each G-tetrad is stabilized by eight hydrogen bonds (four N1–N7 and four O6–N7 interactions) and a central monovalent cation, typically K⁺, which coordinates between the electronegative O6 atoms to enhance folding and thermal stability. These structures are prevalent in telomeric repeats, where they protect chromosome ends from degradation, and in promoter regions of oncogenes like c-MYC, where they regulate transcription by impeding progression. Parallel or antiparallel topologies yield or basket-like folds, with stability increasing with additional stacked tetrads (up to four in a quadruplex). Branched DNA structures, including Y-junctions and cruciforms, further illustrate multimeric assemblies. Y-junctions feature three duplex arms meeting at a central , mimicking replication forks or repair intermediates, with the junction stabilized by continuous base stacking across arms. Four-way junctions, akin to Holliday structures, can form transiently in synthetic or palindromic sequences, exhibiting open or stacked states depending on ion conditions. Cruciforms emerge from sequences under negative supercoiling, extruding two arms from a central four-way junction, which facilitates site-specific recognition in regulatory contexts. In RNA, multimers often involve symmetric dimerization critical for viral genomes. Kissing loops mediate dimerization in retroviral RNAs, such as HIV-1, where complementary loop sequences from two stem-loops base-pair via 6–8 , forming an initial symmetric complex that extends into extended duplexes for packaging. This kissing interaction is highly specific, with stability enhanced by coaxial stacking of adjacent stems, and is conserved across retroviruses to ensure dimer-selective encapsidation. Similarly, small interfering RNAs (siRNAs) form stable 19–21 duplexes with 2-nucleotide overhangs, adopting A-form helices that guide RNA interference; their thermodynamic asymmetry, with weaker 5'-end pairing, facilitates loading into proteins, though the duplex itself is a nucleic acid-only assembly prior to complexation. The stability of these multimers is profoundly influenced by cations. Monovalent ions like K⁺ are essential for G-quadruplexes, occupying the central channel and increasing melting temperatures by up to 20–30°C compared to Na⁺, due to optimal . In contrast, divalent Mg²⁺ ions preferentially stabilize Holliday junctions and branched structures by binding at the backbone near the , screening electrostatic repulsion and promoting the X-conformation with constants in the micromolar range. For kissing loops and siRNA duplexes, Mg²⁺ further enhances loop-loop interactions by reducing charge repulsion, though monovalent ions suffice for basic duplex stability.

Interactions with proteins and other molecules

Nucleic acids frequently form quaternary complexes with proteins, enabling essential cellular processes such as DNA packaging, RNA processing, and gene regulation. In DNA, the serves as the primary unit of organization, where approximately 147 base pairs of DNA wrap around a composed of two copies each of H2A, H2B, , and H4, forming 1.65 left-handed superhelical turns that compact the and facilitate access for regulatory proteins. This wrapping introduces negative supercoiling, which is modulated by histone interactions to influence dynamics. Beyond individual nucleosomes, adopts higher-order folding structures, such as the 30 nm fiber, where arrays of nucleosomes stack via histone H1-mediated interactions, further condensing DNA into looped domains that partition the into functional compartments affecting transcription and replication.00188-2) These protein-DNA assemblies ensure genomic stability while allowing dynamic remodeling by chromatin-modifying enzymes. In RNA, protein interactions are exemplified by ribonucleoprotein complexes like the , a assembly critical for protein synthesis. The bacterial ribosome consists of a small 30S subunit (containing 16S rRNA and 21 proteins) and a large 50S subunit (with 23S and 5S rRNAs and 34 proteins), which together form the 70S ; the rRNAs provide the catalytic core for formation, while proteins stabilize the structure and facilitate mRNA decoding and tRNA binding.00619-0) Similarly, the , responsible for pre-mRNA removal, is a dynamic ribonucleoprotein machine composed of five small nuclear RNPs (snRNPs: U1, , U4, U5, U6) each harboring snRNAs bound to Sm and other proteins, totaling over 300 protein components that assemble stepwise on the to execute splicing catalysis.00146-9) These RNA-protein structures highlight how proteins scaffold RNA folding and enhance functional specificity. Additional interactions involve transcription factors binding to specific DNA motifs, often recognizing sequences through insertion of alpha-helices into the major groove for sequence-specific contacts with base edges and backbones, as seen in motifs of homeodomain proteins.00392-6) In RNA, non-protein ligands can bind structured domains, such as in riboswitches where the (TPP) folds into a three-helix junction that clamps the ligand via hydrogen bonding and base stacking, regulating allosterically without protein intermediaries.00157-0) Binding modes differ between DNA and RNA: DNA-protein recognition primarily exploits the wider major groove for direct readout of base pairs, enabling high specificity, whereas RNA-protein interfaces rely more on shape complementarity, with proteins molding to RNA's irregular 3D surfaces via electrostatic and van der Waals interactions to accommodate diverse folds.

References

  1. [1]
    Nucleic Acids - National Human Genome Research Institute
    Nucleic acids are made of nitrogen-containing bases, phosphate groups, and sugar molecules. Each type of nucleic acid has a distinctive structure and plays a ...
  2. [2]
    Understanding biochemistry: structure and function of nucleic acids
    Oct 11, 2019 · We explain the structure of the DNA molecule, how it is packaged into chromosomes and how it is replicated prior to cell division.
  3. [3]
    A Structure for Deoxyribose Nucleic Acid - Nature
    Molecular Structure of Nucleic Acids: A Structure for Deoxyribose Nucleic Acid. J. D. WATSON &; F. H. C. CRICK. Nature volume 171, pages 737–738 ( ...
  4. [4]
    Role of pKa of Nucleobases in the Origins of Chemical Evolution
    Apr 26, 2012 · Another instructive example is the strong base pairing between two purines, guanine (pKa 9.5) and isoguanine (pKa 9.2) in aqueous medium.
  5. [5]
    Comprehensive Study of the Effects of Methylation on Tautomeric ...
    A comprehensive theoretical study of tautomers of methylated derivatives of guanine, adenine, cytosine, thymine, and uracil was performed.
  6. [6]
    Keeping Uracil Out of DNA: Physiological Role, Structure and ...
    The 5-methyl group, that is, the difference between the two bases, has no effect on the interaction with adenine. In fact, DNA synthesized in a thymine-less ...
  7. [7]
    Complementary Base Pairing: Hydrogen Bonding
    AT and GC pairs maximize the number of hydrogen bonds across the shared helical axis. A's hydrogen donors can pair up with T's hydrogen bond acceptors.
  8. [8]
    Inosine in Biology and Disease - PMC - PubMed Central
    Apr 19, 2021 · Inosine was one of the first nucleobase modifications discovered in nucleic acids, having been identified in 1965 as a component of the first ...
  9. [9]
    [PDF] Chapter 28: Nucleosides, Nucleotides, and Nucleic Acids.
    The chemical linkage between nucleotide units in nucleic acids is a phosphodiester, which connects the 5'-hydroxyl group of one nucleotide to the 3'-hydroxyl ...
  10. [10]
    Chapter 27
    Deoxynucleoside triphosphates (dNTPs) and nucleoside triphosphates (NTPs) serve as the immediate substrates for the biosynthesis of DNA and RNA, respectively ( ...
  11. [11]
    Nucleotide Metabolism - PMC - NIH
    The activity of the NDP kinases is higher than that of the NMP kinases to maintain a high intracellular level of (d)NTPs relative to that of (d)NDPs.
  12. [12]
    Biochemistry, RNA Polymerase - StatPearls - NCBI Bookshelf
    RNA polymerase (RNAP) is a large enzyme that transcribes DNA into RNA. Eukaryotes have three types of RNAPs, while bacteria and archaea have one.
  13. [13]
    New insights into DNA polymerase mechanisms provided by time ...
    DNA polymerases carry out template-directed nucleotide incorporation using the same basic reaction scheme (Fig. 1) [3] . First, the polymerase binds DNA to form ...
  14. [14]
    General Features of DNA Replication - UT Health San Antonio
    Nucleotides are added in a 5'----->3' direction, with anti-parallel complementarity to the sequence of each parental strand, which is termed the template. DNA ...
  15. [15]
    A model for the evolution of nucleotide polymerase directionality
    Apr 22, 2011 · This results in the well known 5'→3' directionality of all DNA and RNA Polymerases. The lack of any alternative mechanism, e.g. addition in a 3' ...
  16. [16]
    Pyrophosphate hydrolysis is an intrinsic and critical step of the DNA ...
    May 30, 2018 · Using time-resolved crystallography, we show that hydrolysis of PPi is an intrinsic and critical step of the DNA synthesis reaction catalyzed by dPols.
  17. [17]
    The Balancing Act of Ribonucleotides in DNA - PMC - NIH
    May 1, 2017 · For DNA Pol α, the rate of incorporation of the incorrect base is about 10−4. DNA Pol δ and ε have the ability to proofread the last ...
  18. [18]
    DNA polymerase α (RNA primase) error rate - Eukaryotes
    High nucleotide selectivity at the polymerase active site is illustrated by the relatively low base substitution and indel error rates (~10^–4) of polymerase a, ...
  19. [19]
    RNA polymerase errors cause splicing defects and can be regulated ...
    Dec 10, 2015 · The experiments show that altering the levels of the different subunits of RNA polymerase in cells can change how many mistakes are made during transcription.
  20. [20]
    Nucleic Acids – Molecular Ecology & Evolution: An Introduction
    The primary structure of DNA refers to the specific sequence of nucleotides. These nucleotides are linked together through covalent bonds, forming a long chain ...
  21. [21]
    Nucleic Acid Structure
    Primary Structure: the sequence of bases along the pentose-phosphodiester backbone of a DNA molecule (or an RNA molecule) · 5 end to the 3 end--directionality ...
  22. [22]
    How to Calculate DNA Melting Temperature (Tm) | IDT
    Apr 17, 2023 · Learn how to calculate DNA melting temperature (Tm) using proven formulas to improve primer specificity and PCR performance.
  23. [23]
    Chargaff's Rules: the Work of Erwin Chargaff
    Oct 9, 2020 · His immediate challenge was to devise a method toanalyze the nitrogenous components and sugars of DNA from different species.Because large ...<|separator|>
  24. [24]
    3.12: DNA and RNA - Biology LibreTexts
    Jul 28, 2023 · Rule 1. Chargaff determined that in DNA, the amount of one base, a purine, always approximately equals the amount of a particular second base, ...Discovery that DNA is the... · Chargaff Focuses on DNA Bases · DNA Replication
  25. [25]
    Roles of mRNA poly(A) tails in regulation of eukaryotic gene ...
    Mar 13, 2023 · In a general model of canonical mRNA decay, the poly(A) tail is first shortened to 10–12 nt, then the 5′-cap is removed (decapping) and the ...
  26. [26]
    DNA Methylation and Its Basic Function | Neuropsychopharmacology
    Jul 11, 2012 · DNA methylation regulates gene expression by recruiting proteins involved in gene repression or by inhibiting the binding of transcription factor(s) to DNA.
  27. [27]
    Mapping epigenetic modifications by sequencing technologies
    Sep 1, 2023 · DNA methylation of cytosine (5mC) is the most predominant modification, and accounts for ~5% of all cytosine (C) [42]. 5mC modification on DNA ...
  28. [28]
    N6-Methyladenine DNA modification in Xanthomonas oryzae pv ...
    Nov 2, 2018 · DNA N6-methyladenine (6mA) modifications expand the information capacity of DNA and have long been known to exist in bacterial genomes.
  29. [29]
    The role of m6A, m5C and Ψ RNA modifications in cancer
    Jan 18, 2021 · In this Review, we focus on the occurrence of N 6 -methyladenosine (m 6 A), 5-methylcytosine (m 5 C) and pseudouridine (Ψ) in coding and non-coding RNAs.
  30. [30]
    Pseudouridine Formation, the Most Common Transglycosylation in ...
    Pseudouridine (Ψ) synthases are the enzymes responsible for the most abundant and phylogenetically conserved posttranscriptional modification of cellular RNAs.
  31. [31]
    Therapeutic siRNA: state of the art - Nature
    Jun 19, 2020 · Early siRNA therapeutic programs typically employ partial or slight modifications by using 2′-OMe, 2′-F or PS, e.g., alternative 2′-OMe ...
  32. [32]
    Interaction preferences between protein side chains and key ...
    Nov 15, 2022 · The study showed that 5mC has potential to reduce the flexibility of DNA ... Epigenetics of modified DNA bases: 5-methylcytosine and beyond.
  33. [33]
    Unique Gene‐Silencing and Structural Properties of 2′‐Fluoro ...
    Jan 31, 2011 · We also found that siRNAs modified with 2′-F exhibited increased nuclease stability, significantly decreased immune stimulation in an in ...
  34. [34]
    Ultrafast bisulfite sequencing detection of 5-methylcytosine in DNA ...
    Jan 2, 2024 · 5-Methylcytosine (5mC) in DNA is a fundamental epigenetic mark having crucial roles in regulating gene expression, chromatin organization and ...
  35. [35]
    The DNA Base Pairs in Detail - LabXchange
    Adenine-thymine base pairs, or AT base pairs, are held together by 2 hydrogen bonds. Guanine-cytosine base pairs, or GC base pairs, are held together by 3 ...
  36. [36]
    Cooperative Hydrogen-Bonding in Adenine−Thymine and Guanine ...
    The A·T base pair has two traditional H-bonds and one weak C−H···O interaction. After BSSE correction, the interaction energies are calculated to be −11.7 and ...Methods · Results and Discussion · Conclusions · References
  37. [37]
    Base Pair - National Human Genome Research Institute
    The two strands are held together by hydrogen bonds between pairs of bases: adenine pairs with thymine, and cytosine pairs with guanine.
  38. [38]
    The RNA Base Pairs - LabXchange
    Sep 7, 2021 · Each base pair consists of 2 nucleotides held together by hydrogen bonds. The core of each nucleotide is a pentose whose carbon residues are numbered 1 prime ...
  39. [39]
    Estimating Strengths of Individual Hydrogen Bonds in RNA Base Pairs
    Apr 23, 2019 · Each base pair is composed of at least two interbase hydrogen bonds (H-bonds). It is expected that the characteristic geometry and stability of ...Introduction · Methods · Results and Discussion · Supporting Information
  40. [40]
    Codon—anticodon pairing: The wobble hypothesis - ScienceDirect
    This hypothesis is explored systematically, and it is shown that such a wobble could explain the general nature of the degeneracy of the genetic code.
  41. [41]
    The G·U wobble base pair: A fundamental building block of RNA ...
    Watson–Crick G·C and A·U base pairs differ from wobble G·U pairs in the type and location of functional groups that are projected into major and minor grooves.
  42. [42]
    Optimised parameters for A-DNA and B-DNA - PubMed
    Optimised parameters for A-DNA and B-DNA. ... Authors. S Arnott, D W L Hukins. PMID: 5040245; DOI: 10.1016/0006-291X(72)90243 ...
  43. [43]
    RNA bulges as architectural and recognition motifs - ScienceDirect
    Bulges are unpaired stretches of nucleotides located within one strand of a nucleic acid duplex which is formed by hydrogen-bonded bases.
  44. [44]
    Stability of single-nucleotide bulge loops embedded in a GAAA RNA ...
    One of the more common secondary structural motifs, a bulge loop, occurs when a duplex is interrupted by one or more unpaired nucleotides in one of the strands.
  45. [45]
    Structural Features that Give Rise to the Unusual Stability of RNA ...
    The most frequently occurring RNA hairpins in 16Sand 23Sribosomal RNA contain a tetranucleotide loop that has a GNRA consensus sequence.Missing: seminal | Show results with:seminal
  46. [46]
    Thermodynamic parameters for loop formation in RNA and DNA ...
    We have determined the contributions of tetraloops to the thermodynamic stability of RNA and DNA hairpins. Included in this study are those loops which have ...Missing: seminal paper
  47. [47]
    Pseudoknots: RNA Structures with Diverse Functions | PLOS Biology
    Jun 14, 2005 · A pseudoknot is an RNA structure that is minimally composed of two helical segments connected by single-stranded regions or loops.
  48. [48]
    A combinatorial approach to the repertoire of RNA kissing motifs
    Apr 11, 2016 · Loop–loop (also known as kissing) interactions between RNA hairpins are involved in several mechanisms in both prokaryotes and eukaryotes ...
  49. [49]
    tRNA‐like structures and their functions - Wu - 2022 - FEBS Press
    Jun 12, 2021 · The cloverleaf structure consists of an acceptor stem, a D-Arm, an anticodon stem-loop, and a T-Arm. It folds into a L-shaped three-dimensional ...
  50. [50]
    New RNA motifs suggest an expanded scope for riboswitches in ...
    These findings indicate that riboswitches serve as a major genetic regulatory mechanism for the control of metabolic genes in many microbial species.
  51. [51]
    Mfold web server for nucleic acid folding and hybridization prediction
    The original applications on the mfold web server deal with folding a single RNA or DNA sequence per submission (job). The submission forms for RNA and DNA are ...Abstract · INTRODUCTION · SERVER CONTENT AND... · FUTURE DIRECTIONS
  52. [52]
    ViennaRNA Package 2.0 | Algorithms for Molecular Biology | Full Text
    Nov 24, 2011 · The main secondary structure prediction tool is RNAfold, which computes the minimum free energy (MFE) and backtraces an optimal secondary ...
  53. [53]
    Topological challenges to DNA replication: Conformations at the fork
    Lk is the sum of Tw and Wr. Notably, Lk is unaltered by any deformation short of DNA breakage and reunion. More important is the quantity ΔLk, the difference ...
  54. [54]
    The Writhing Number of a Space Curve - PNAS
    The writhing number measures the extent to which coiling of the central curve has relieved local twisting of the cord.
  55. [55]
    DNA supercoiling in bacteria: state of play and challenges from a ...
    This fundamental characteristic enables supercoiled DNA to relieve local torsional stress by generating super-structures, such as plectonemes. The relative ...
  56. [56]
    Interaction between DNA and an Escherichia coli protein omega
    Interaction between DNA and an Escherichia coli protein omega. J Mol Biol. 1971 Feb 14;55(3):523-33. doi: 10.1016 ...Missing: topoisomerase IJ
  57. [57]
    DNA Gyrase and the Supercoiling of DNA - Science
    Negative supercoiling of bacterial DNA by DNA gyrase influences all metabolic processes involving DNA and is essential for replication.
  58. [58]
    DNA topology in chromatin is defined by nucleosome spacing
    Oct 27, 2017 · We demonstrate that the DNA linking number per nucleosome strongly depends on the nucleosome spacing and varies from −1.4 to −0.9.
  59. [59]
    DNA supercoiling and transcription in bacteria: a two-way street
    Jul 18, 2019 · Type II topoisomerases help to resolve the DNA topological conflicts ... topoisomerase I, is stimulated by negative DNA supercoiling [67, 68].<|separator|>
  60. [60]
    DNA supercoiling is a fundamental regulatory principle in the control ...
    DNA supercoiling, arising from underwinding, influences gene expression by impacting transcription and connecting DNA topology to gene expression.
  61. [61]
    Computational modeling of RNA 3D structures, with the aid of ... - NIH
    In this work, we review computational approaches for RNA structure prediction, with emphasis on implementations (particular programs) that can utilize ...Missing: paper | Show results with:paper
  62. [62]
    An RNA-centric historical narrative around the Protein Data Bank
    ... stems, the acceptor-stem with the Thymine (T)-stem and of the Dihydrouridine (D)-stem with the anticodon stem. A stem capped by a loop is called a hairpin.
  63. [63]
    The diverse structural modes of tRNA binding and recognition - NIH
    All tRNAs fold into a relatively rigid three-dimensional L-shaped structure. The conserved tertiary organization of canonical tRNA arises through the formation ...Missing: coaxial seminal
  64. [64]
    [PDF] The Structural Basis of Hammerhead Ribozyme Self-Cleavage
    Mar 6, 1998 · This ribozyme construct employs a 16-nucleotide enzyme strand (green) and a 25-nucleotide substrate strand (blue and red). The letters outlined ...
  65. [65]
    Structural basis for the fast self-cleavage reaction catalyzed ... - PNAS
    The small self-cleaving ribozyme family can be split into two groups based on whether their active site is formed by an irregular helix (hammerhead, hairpin, ...
  66. [66]
    Crystal structure of a group I intron splicing intermediate - PMC
    A recently reported crystal structure of an intact bacterial group I self-splicing intron in complex with both its exons provided the first molecular view ...
  67. [67]
    Recognition modes of RNA tetraloops and tetraloop-like motifs by ...
    RNA tetraloops can mediate RNA-RNA contacts via the tetraloop-receptor motif, kissing hairpin loops, A-minor interactions, and pseudoknots.Missing: 3D packing seminal paper
  68. [68]
    RNA Structural Dynamics As Captured by Molecular Simulations
    Jan 3, 2018 · Our review encompasses tetranucleotides, tetraloops, a number of small RNA motifs, A-helix RNA, kissing-loop complexes, the TAR RNA element, the ...
  69. [69]
    Modeling Complex RNA Tertiary Folds with Rosetta - ScienceDirect
    In this chapter, we give a practical guide to our current workflow for modeling RNA three-dimensional structures using FARFAR.Missing: paper | Show results with:paper
  70. [70]
    Chromatin fibers stabilize nucleosomes under torsional stress - Nature
    Jan 8, 2020 · The nucleosomal DNA winds around a histone octamer 1.65 times in a left-handed manner and has a slightly lower helical twist density than bare ...
  71. [71]
    Intracellular nucleosomes constrain a DNA linking number ... - Nature
    Sep 28, 2018 · Nucleosomal Wr is about −1.53 when the core DNA completes 1.65 left-handed superhelical turns around the histone octamer. This conformation ...
  72. [72]
    RNA structure drives interaction with proteins - Nature
    Jul 19, 2019 · In this large spectrum of activities, RNA structure controls the precise binding of proteins by creating spatial patterns and alternative ...