首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
It was established that the cytosine·thymine (C·T) mismatched DNA base pair with cis‐oriented N1H glycosidic bonds has propeller‐like structure (|N3C4C4N3| = 38.4°), which is stabilized by three specific intermolecular interactions–two antiparallel N4H…O4 (5.19 kcal mol?1) and N3H…N3 (6.33 kcal mol?1) H‐bonds and a van der Waals (vdW) contact O2…O2 (0.32 kcal mol?1). The C·T base mispair is thermodynamically stable structure (ΔGint = ?1.54 kcal mol?1) and even slightly more stable than the A·T Watson–Crick DNA base pair (ΔGint = ?1.43 kcal mol?1) at the room temperature. It was shown that the C·T ? C*·T* tautomerization via the double proton transfer (DPT) is assisted by the O2…O2 vdW contact along the entire range of the intrinsic reaction coordinate (IRC). The positive value of the Grunenberg's compliance constants (31.186, 30.265, and 22.166 Å/mdyn for the C·T, C*·T*, and TSC·T ? C*·T*, respectively) proves that the O2…O2 vdW contact is a stabilizing interaction. Based on the sweeps of the H‐bond energies, it was found that the N4H…O4/O4H…N4, and N3H…N3 H‐bonds in the C·T and C*·T* base pairs are anticooperative and weaken each other, whereas the middle N3H…N3 H‐bond and the O2…O2 vdW contact are cooperative and mutually reinforce each other. It was found that the tautomerization of the C·T base mispair through the DPT is concerted and asynchronous reaction that proceeds via the TSC·T ? C*·T* stabilized by the loosened N4? H? O4 covalent bridge, N3H…N3 H‐bond (9.67 kcal mol?1) and O2…O2 vdW contact (0.41 kcal mol?1). The nine key points, describing the evolution of the C·T ? C*·T* tautomerization via the DPT, were detected and completely investigated along the IRC. The C*·T* mispair was revealed to be the dynamically unstable structure with a lifetime 2.13·× 10?13 s. In this case, as for the A·T Watson–Crick DNA base pair, activates the mechanism of the quantum protection of the C·T DNA base mispair from its spontaneous mutagenic tautomerization through the DPT. © 2013 Wiley Periodicals, Inc.  相似文献   

2.
Ultrafast deactivation pathways bestow photostability on nucleobases and hence preserve the structural integrity of DNA following absorption of ultraviolet (UV) radiation. One controversial recovery mechanism proposed to account for this photostability involves electron‐driven proton transfer (EDPT) in Watson–Crick base pairs. The first direct observation is reported of the EDPT process after UV excitation of individual guanine–cytosine (G?C) Watson–Crick base pairs by ultrafast time‐resolved UV/visible and mid‐infrared spectroscopy. The formation of an intermediate biradical species (G[?H]?C[+H]) with a lifetime of 2.9 ps was tracked. The majority of these biradicals return to the original G?C Watson–Crick pairs, but up to 10 % of the initially excited molecules instead form a stable photoproduct G*?C* that has undergone double hydrogen‐atom transfer. The observation of these sequential EDPT mechanisms across intermolecular hydrogen bonds confirms an important and long debated pathway for the deactivation of photoexcited base pairs, with possible implications for the UV photochemistry of DNA.  相似文献   

3.
Janus bases are heterocyclic nucleic acid base analogs that present two different faces able to simultaneously hydrogen bond to nucleosides that form Watson–Crick base pairs. The synthesis of a Janus‐AT nucleotide analogue, N JAT , that has an additional endocyclic ring nitrogen and is thus more capable of efficiently discriminating T/A over G/C bases when base‐pairing in a standard duplex‐DNA context is described. Conversion to a phosphoramidite ultimately afforded incorporation into an oligonucleotide. In contrast to the first generation of carbocyclic Janus heterocycles, it remains in its unprotonated state at physiological pH and, therefore, forms very stable Watson–Crick base pairs with either A or T bases. Biophysical and computational methods indicate that N JAT is an improved candidate for sequence‐specific genome targeting.  相似文献   

4.
Reverse Watson–Crick DNA with parallel‐strand orientation (ps DNA) has been constructed. Pyrrolo‐dC (PyrdC) nucleosides with phenyl and pyridinyl residues linked to the 6 position of the pyrrolo[2,3‐d]pyrimidine base have been incorporated in 12‐ and 25‐mer oligonucleotide duplexes and utilized as silver‐ion binding sites. Thermal‐stability studies on the parallel DNA strands demonstrated extremely strong silver‐ion binding and strongly enhanced duplex stability. Stoichiometric UV and fluorescence titration experiments verified that a single 2pyPyrdC–2pyPyrdC pair captures two silver ions in ps DNA. A structure for the PyrdC silver‐ion base pair that aligns 7‐deazapurine bases head‐to‐tail instead of head‐to‐head, as suggested for canonical DNA, is proposed. The silver DNA double helix represents the first example of a ps DNA structure built up of bidentate and tridentate reverse Watson–Crick base pairs stabilized by a dinuclear silver‐mediated PyrdC pair.  相似文献   

5.
Self‐assembled systems rely on intramolecular cooperative effects to control their growth and regulate their shape, thus yielding discrete, well‐defined structures. However, as the size of the system increases, cooperative effects tend to dissipate. We analyze here this situation by studying a set of oligomers of different lengths capped with guanosine and cytidine nucleosides, which associate in cyclic tetramers by complementary Watson–Crick H‐bonding interactions. As the monomer length increases, and thus the number of C(sp)–C(sp2) σ ‐bonds in the π‐conjugated skeleton, the macrocycle stability decreases due to a notable reduction in effective molarity (EM), which has a clear entropic origin. We determined the relationship between EM or ΔS and the number of σ‐bonds, which allowed us to predict the maximum monomer lengths at which cyclic species would still assemble quantitatively, or whether the cyclic species would not able to compete at all with linear oligomers over the whole concentration range.  相似文献   

6.
The title mononuclear complex, [Cu(CO3)(C8H7N5)(H2O)]·2H2O, was obtained by fixation of CO2 by a mixture of copper(II) tetra­fluoro­borate and the ligand bis­(pyrimidin‐2‐yl)­amine in ethanol/water. The CuII ion of the complex has a distorted square‐pyramidal environment, with a basal plane formed by two N atoms of the ligand and two chelating O atoms of the carbonate group, while the apical position is occupied by the O atom of the coordinating water mol­ecule. In the solid state, hydrogen‐bonding interactions are dominant, the most unusual being the Watson–Crick‐type coplanar ligand pairing through two N—H?N bonds. Lattice water mol­ecules also participate in hydrogen bonding.  相似文献   

7.
Two novel cocrystals of the N(7)—H tautomeric form of N6‐benzoyladenine (BA), namely N6‐benzoyladenine–3‐hydroxypyridinium‐2‐carboxylate (3HPA) (1/1), C12H9N5O·C6H5NO3, (I), and N6‐benzoyladenine–DL‐tartaric acid (TA) (1/1), C12H9N5O·C4H6O6, (II), are reported. In both cocrystals, the N6‐benzoyladenine molecule exists as the N(7)—H tautomer, and this tautomeric form is stabilized by intramolecular N—H...O hydrogen bonding between the benzoyl C=O group and the N(7)—H hydrogen on the Hoogsteen site of the purine ring, forming an S(7) motif. The dihedral angle between the adenine and phenyl planes is 0.94 (8)° in (I) and 9.77 (8)° in (II). In (I), the Watson–Crick face of BA (N6—H and N1; purine numbering) interacts with the carboxylate and phenol groups of 3HPA through N—H...O and O—H...N hydrogen bonds, generating a ring‐motif heterosynthon [graph set R22(6)]. However, in (II), the Hoogsteen face of BA (benzoyl O atom and N7; purine numbering) interacts with TA (hydroxy and carbonyl O atoms) through N—H...O and O—H...O hydrogen bonds, generating a different heterosynthon [graph set R22(4)]. Both crystal structures are further stabilized by π–π stacking interactions.  相似文献   

8.
Substituted Watson–Crick guanine–cytosine (GC) base pairs were recently shown to yield robust three‐state nanoswitches. Here, we address the question: Can such supramolecular switches also be based on Watson–Crick adenine‐thymine (AT) base pairs? We have theoretically analyzed AT pairs in which purine‐C8 and/or pyrimidine‐C6 positions carry a substituent X=NH?, NH2, NH3+ (N series), O?, OH or OH2+ (O series), using the generalized gradient approximation (GGA) of density functional theory at the BP86/TZ2P level. Thus, we explore the trend in geometrical shape and hydrogen bond strengths in AT pairs along a series of stepwise protonations of the substituents. Introducing a charge on the substituents leads to substantial and characteristic changes in the individual hydrogen bond lengths when compared to the neutral AT pair. However, the trends along the series of negative, neutral, and positive substituents are less systematic and less pronounced than for GC. In certain instances, internal proton transfer from thymine to adenine occurs. Our results suggest that AT is a less suitable candidate than GC in the quest for chemically controlled nanoswitches.  相似文献   

9.
Metallo‐base pairs have been extensively studied for applications in nucleic acid‐based nanodevices and genetic code expansion. Metallo‐base pairs composed of natural nucleobases are attractive because nanodevices containing natural metallo‐base pairs can be easily prepared from commercially available sources. Previously, we have reported a crystal structure of a DNA duplex containing T? HgII? T base pairs. Herein, we have determined a high‐resolution crystal structure of the second natural metallo‐base pair between pyrimidine bases C? AgI? C formed in an RNA duplex. One AgI occupies the center between two cytosines and forms a C? AgI? C base pair through N3? AgI? N3 linear coordination. The C? AgI? C base pair formation does not disturb the standard A‐form conformation of RNA. Since the C? AgI? C base pair is structurally similar to the canonical Watson–Crick base pairs, it can be a useful building block for structure‐based design and fabrication of nucleic acid‐based nanodevices.  相似文献   

10.
Cardiosulfa is a biologically active sulfonamide molecule that was recently shown to induce abnormal heart development in zebrafish embryos through activation of the aryl hydrocarbon receptor (AhR). The present report is a systematic study of solid‐state forms of cardiosulfa and its biologically active analogues that belong to the N‐(9‐ethyl‐9H‐carbazol‐3‐yl)benzene sulfonamide skeleton. Cardiosulfa (molecule 1 ; R1=NO2, R2=H, R3=CF3), molecule 2 (H, H, CF3), molecule 3 (CF3, H, H), molecule 4 (NO2, H, H), molecule 5 (H, CF3, H), and molecule 6 (H, H, H) were synthesized and subjected to a polymorph search and solid‐state form characterization by X‐ray diffraction, differential scanning calorimetry (DSC), variable‐temperature powder X‐ray diffraction (VT‐PXRD), FTIR, and solid‐state (ss) NMR spectroscopy. Molecule 1 was obtained in a single‐crystalline modification that is sustained by N? H???π and C? H???O interactions but devoid of strong intermolecular N? H???O hydrogen bonds. Molecule 2 displayed a N? H???O catemer C(4) chain in form I, whereas a second polymorph was characterized by PXRD. The dimorphs of molecule 3 contain N? H???π and C? H???O interactions but no N? H???O bonds. Molecule 4 is trimorphic with N? H???O catemer in form I, and N? H???π and C? H???O interactions in form II, and a third polymorph was characterized by PXRD. Both polymorphs of molecule 5 contain the N? H???O catemer C(4) chain, whereas the sulfonamide N? H???O dimer synthon R22(8) was observed in polymorphs of 6 . Differences in the strong and weak hydrogen‐bond motifs were correlated with the substituent groups, and the solubility and dissolution rates were correlated with the conformation in the crystal structure of 1 , 2 , 3 , 4 , 5 , 6 . Higher solubility compounds, such as 2 (10.5 mg mL?1) and 5 (4.4 mg mL?1), adopt a twisted confirmation, whereas less‐soluble 1 (0.9 mg mL?1) is nearly planar. This study provides practical guides for functional‐group modification of drug lead compounds for solubility optimization.  相似文献   

11.
Incorporation of deoxynucleotide analogues into DNA is important for the expansion of DNA functions. Primer extension reactions are commonly used for the assay of such reaction events. However, current assay protocols generally rely on radiolabeling, fluorescence reporter labeling, or removal of specific deoxynucleotide triphosphate in the reaction mixture. Herein we report on the design of two novel assay protocols that utilize a dideoxynucleotide‐terminated template strand and a phosphorothiolate‐modified deoxynucleotide‐terminated template strand. We designed and synthesized a deoxyuridine triphosphate analogue (dU*TP) containing 2‐bromoisobutyryl group and demonstrated that it could be well recognized by ?29DNA polymerase, E. coli DNA polymerase I Klenow Fragment, Bst DNA polymerase Large Fragment, and E. coli DNA polymerase I Klenow Fragment (exo(), which translated to effective incorporation of dU*TP into DNA. dU*TP was also successfully incorporated into extremely long single‐stranded DNA at high‐density using ?29 DNA polymerase by rolling circle amplification.  相似文献   

12.
The self‐complementary UA and AU dinucleotide analogues 41 – 45, 47, 48 , and 51 – 60 were prepared by Sonogashira coupling of 6‐iodouridines with C(5′)‐ethynylated adenosines and of 8‐iodoadenosines with C(5′)‐ethynylated uridines. The dinucleotide analogues associate in CDCl3 solution. The C(6/I)‐unsubstituted AU dimers 51 and 54 prefer an anti‐oriented uracilyl group and form stretched linear duplexes. The UA propargyl alcohols 41 and 43 – 45 possess a persistent intramolecular O(5′/I)? H???N(3/I) H‐bond and, thus, a syn‐oriented adeninyl and a gt‐ or tg‐oriented ethynyl moiety; they form corrugated linear duplexes. All other dimers form cyclic duplexes characterized by syn‐oriented nucleobases. The preferred orientation of the ethynyl moiety (the C(4′),C(5′) torsion angle) defines a conformation between gg and one where the ethynyl group eclipses O(4′/I). The UA dimers 42, 47 , and 48 form Watson–Crick H‐bonds, the AU dimers 56 and 58 – 60 H‐bonds of the Watson–Crick‐type, the AU dimers 53 and 55 reverse‐Hoogsteen, and 57 Hoogsteen H‐bonds. The pairing mode depends on the substituent of C(5′/I) (H, OSiiPr3; OH) and on the H‐bonds of HO? C(5′/I) in the AU dimers. Association constants were derived from the concentration‐dependent chemical shift for HN(3) of the uracilyl moiety; they vary from 45–104 M ?1 for linear duplexes to 197–2307 M ?1 for cyclic duplexes. The thermodynamic parameters were determined by van't Hoff analysis of the temperature‐dependence of the (concentration‐dependent) chemical shift for HN(3) of the uracilyl moiety. Neglecting stacking energies, one finds an average energy of 3.5–4.0 kcal/mol per intermolecular H‐bond. Base stacking is evidenced by the temperature‐dependent CD spectra. The crystal structure of 54 shows two antiparallel chains of dimers connected by Watson‐Crick H‐bonds. The chains are bridged by a strong H‐bond between the propargylic OH and O?C(4) and by weak reverse A ? A Hoogsteen H‐bonds.  相似文献   

13.
Intramolecular H‐atom transfer in model peptide‐type radicals was investigated with high‐level quantum‐chemistry calculations. Examination of 1,2‐, 1,3‐, 1,5‐, and 1,6[C ? N]‐H shifts, 1,4‐ and 1,7[C ? C]‐H shifts, and 1,4[N ? N]‐H shifts (Scheme 1), was carried out with a number of theoretical methods. In the first place, the performance of UB3‐LYP (with the 6‐31G(d), 6‐31G(2df,p), and 6‐311+G(d,p) basis sets) and UMP2 (with the 6‐31G(d) basis set) was assessed for the determination of radical geometries. We found that there is only a small basis‐set dependence for the UB3‐LYP structures, and geometries optimized with UB3‐LYP/6‐31G(d) are generally sufficient for use in conjunction with high‐level composite methods in the determination of improved H‐transfer thermochemistry. Methods assessed in this regard include the high‐level composite methods, G3(MP2)‐RAD, CBS‐QB3, and G3//B3‐LYP, as well as the density‐functional methods B3‐LYP, MPWB1K, and BMK in association with the 6‐31+G(d,p) and 6‐311++G(3df,3pd) basis sets. The high‐level methods give results that are close to one another, while the recently developed functionals MPWB1K and BMK provide cost‐effective alternatives. For the systems considered, the transformation of an N‐centered radical to a C‐centered radical is always exothermic (by 25 kJ ? mol?1 or more), and this can lead to quite modest barrier heights of less than 60 kJ ? mol?1 (specifically for 1,5[C ? N]‐H and 1,6[C ? N]‐H shifts). H‐Migration barriers appear to decrease as the ring size in the transition structure (TS) increases, with a lowering of the barrier being found, for example when moving from a rearrangement proceeding via a four‐membered‐ring TS (e.g., the 1,3[C ? N]‐H shift, CH3? C(O)? NH..CH2? C(O)? NH2) to a rearrangement proceeding via a six‐membered‐ring TS (e.g., the 1,5[C ? N]‐H shift, .NH? CH2? C(O)? NH? CH3 → NH2? CH2? C(O)? NH? CH2.).  相似文献   

14.
The photoelectron spectrum of the anion of the guanine ??? cytosine base apair (GC).? is recorded for the first time. The observed variation in the spectral peak‐height ratios with the source conditions suggests the presence of two or more anionic isomers. Two maxima of the broad bands in the photoelectron spectrum were measured at about 1.9 and about 2.6 eV. These values are very well reproduced by the vertical detachment energies of the B3LYP/6‐31++G(d,p) calculated low‐energy anionic structures, which are 1) the Watson–Crick base‐pair anion with proton transferred from N1 of guanine to N3 of cytosine, 2) its analogue in which the proton is transferred from N9 of guanine to N7 of guanine, and 3) the global minimum geometry, which is formed from the latter anion by rotation of guanine about the axis approximately defined by C2 of guanine and C4 of cytosine. Furthermore, a minor difference in the stabilities of the two lowest energy anions explains the experimentally observed source (temperature) dependence of the PES spectrum. A rational procedure, based on the chemistry involved in the formation of anionic dimers, which enables the low‐energy anions populated in the photoelectron spectrum to be identified is proposed. In contrast to the alternative combinatorial approach, which in the studied case would lead to carrying out quantum chemical calculations for 2000–2500 structures, the procedure described here reduces the computational problem to only 15 geometries.  相似文献   

15.
The 6‐aminopyrazin‐2(1H)‐one, when incorporated as a pyrimidine‐base analog into an oligonucleotide chain, presents a H‐bond donor? donor? acceptor pattern to a complementary DNA or RNA strand. When paired with the corresponding acceptor? acceptor? donor purine in oligonucleotides, the heterocycle selectively contributes to the stability of the duplex, presumably by forming a base pair of Watson? Crick geometry joined by a nonstandard H‐bonding pattern, expanding the genetic alphabet. Reported here is a short, high yielding, β‐D ‐selective synthesis of a 6‐aminopyrazin‐2(1H)‐one nucleoside via the glycine riboside derivative 28 . The key steps include a Wittig? Horner reaction of an appropriately protected ribose derivative (Scheme 10, 19 → 21 ) followed by a Michael‐like ring closure (Scheme 12, 30 → 1a and 32 → 1b ). Thus, a variety of pyrazine nucleosides (Scheme 13) including the target 6‐aminopyrazin‐2(1H)‐one riboside 1a , and its 5‐methyl derivative 1b , 6‐amino‐5‐methylpyrazin‐2(1H)‐one riboside, are obtained.  相似文献   

16.
《Chemphyschem》2003,4(8):838-842
The vibronic spectrum of the adenine–thymine (A–T) base pair was obtained by one‐color resonant two‐photon ionization (R2PI) spectroscopy in a free jet of thermally evaporated A and T under conditions favorable for formation of small clusters. The onset of the spectrum at 35 064 cm?1 exhibits a large red shift relative to the π–π* origin of 9H‐adenine at 36 105 cm?1. The IR–UV spectrum was assigned to cluster structures with HNH???O?C/N???HN hydrogen bonding by comparison with the IR spectra of A and T monomers and with ab initio calculated vibrational spectra of the most stable A–T isomers. The Watson–Crick A–T base pair is not the most stable base‐pair structure at different levels of ab initio theory, and its vibrational spectrum is not in agreement with the observed experimental spectrum. Experiments with methylated A and T were performed to further support the structural assignment.  相似文献   

17.
Recent experimental studies on the Watson–Crick type base pairing of triazine and aminopyrimidine derivatives suggest that acid/base properties of the constituent bases might be related to the duplex stabilities measured in solution. Herein we use high‐level quantum chemical calculations and molecular dynamics simulations to evaluate the base pairing and stacking interactions of seven selected base pairs, which are common in that they are stabilized by two N? H???O hydrogen bonds separated by one N? H???N hydrogen bond. We show that neither the base pairing nor the base stacking interaction energies correlate with the reported pKa data of the bases and the melting points of the duplexes. This suggests that the experimentally observed correlation between the melting point data of the duplexes and the pKa values of the constituent bases is not rooted in the intrinsic base pairing and stacking properties. The physical chemistry origin of the observed experimental correlation thus remains unexplained and requires further investigations. In addition, since our calculations are carried out with extrapolation to the complete basis set of atomic orbitals and with inclusion of higher electron correlation effects, they provide reference data for stacking and base pairing energies of non‐natural bases.  相似文献   

18.
Structural characteristics of Watson–Crick hydrogen-bonded base pairs are displayed by methylene-bridged base pairs of type A . The shown superposition of the X-ray structure obtained for the base pair A (Rib1=Et; Rib2=Me) over that of a C–G base pair illustrates that A occupies an area similar to that occupied by a traditional Watson–Crick hydrogen-bonded base pair. Temperature-dependent 1H NMR studies indicate that the energy barrier for rotation along its CH2 bridge is about 10 kcal mol−1, and that it exists predominantly in one conformer at −70°C.  相似文献   

19.
The oligonucleotide d(TX)9, which consists of an octadecamer sequence with alternating non‐canonical 7‐deazaadenine (X) and canonical thymine (T) as the nucleobases, was synthesized and shown to hybridize into double‐stranded DNA through the formation of hydrogen‐bonded Watson–Crick base pairs. dsDNA with metal‐mediated base pairs was then obtained by selectively replacing W‐C hydrogen bonds by coordination bonds to central silver(I) ions. The oligonucleotide I adopts a duplex structure in the absence of Ag+ ions, and its stability is significantly enhanced in the presence of Ag+ ions while its double‐helix structure is retained. Temperature‐dependent UV spectroscopy, circular dichroism spectroscopy, and ESI mass spectrometry were used to confirm the selective formation of the silver(I)‐mediated base pairs. This strategy could become useful for preparing stable metallo‐DNA‐based nanostructures.  相似文献   

20.
In this study, using QM/QTAIM calculations in the continuum with ε = 1 under normal conditions, we have revealed for the first time the nondissociative A·T(WC)↔A·T(rWC)/A·T(rH) and A·T(H)↔A·T(rH)/A·T(rWC) conformational transitions. It was established that they proceed via the essentially nonplanar transition states (С1 symmetry) through the intermediates, which are wobbled conformers (С1 symmetry) theoretically predicted in our previous work (Brovarets’ et al., Frontiers in Chemistry, 2018, 6:8, 10.3389/fchem.2018.00008) of the classical А·Т DNA base pairs—Watson–Crick А·Т(WC), reverse Watson–Crick А·Т(rWC), Hoogsteen А·Т(Н) and reverse Hoogsteen А·Т(rН). At this, the A·T(H)↔A·T(rWC) and A·T(WC)↔A·T(rH) conformational transformations are controlled by the transition states (TSs) stabilized by the participation of the intermolecular (T)N3H···N6(A) H‐bond (∼3.70 kcal·mol−1) between the imino group N3H of T and pyramidilized amino group N6H2 of A. Gibbs free energies of activation for these processes consist 12.22 and 11.11 kcal·mol−1, accordingly, under normal conditions. TSs, which control the A·T(WC)↔A·T(rWC) and A·T(H)↔A·T(rH) conformational transitions are stabilized by the participation of the intermolecular (T)N3H···N6(A) H‐bond (5.82 kcal·mol−1) and bifurcating intermolecular (T)N3H···N6(A) (5.00) and (T)N3H···N7(A) (0.61 kcal·mol−1) H‐bonds, accordingly. Notably, in these two TSs amino group N6H2 of A is significantly pyramidilized; Gibbs free energies of activation for these reactions are 19.07 and 19.71 kcal·mol−1, accordingly.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号