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 203 毫秒
1.
We have investigated the production of reactive oxygen species (ROS) by Complex I in isolated open bovine heart submitochondrial membrane fragments during forward electron transfer in presence of NADH, by means of the probe 2′,7′-Dichlorodihydrofluorescein diacetate. ROS production by Complex I is strictly related to its inhibited state. Our results indicate that different Complex I inhibitors can be grouped into two classes: Class A inhibitors (Rotenone, Piericidin A and Rolliniastatin 1 and 2) increase ROS production; Class B inhibitors (Stigmatellin, Mucidin, Capsaicin and Coenzyme Q2) prevent ROS production also in the presence of Class A inhibitors. Addition of the hydrophilic Coenzyme Q1 as an electron acceptor potentiates the effect of Rotenone-like inhibitors in increasing ROS production, but has no effect in the presence of Stigmatellin-like inhibitors; the effect is not shared by more hydrophobic quinones such as decyl-ubiquinone. This behaviour relates the prooxidant CoQ1 activity to a hydrophilic electron escape site. Moreover the two classes of Complex I inhibitors have an opposite effect on the increase of NADH-DCIP reduction induced by short chain quinones: only Class B inhibitors allow this increase, indicating the presence of a Rotenone-sensitive but Stigmatellin-insensitive semiquinone species in the active site of the enzyme. The presence of this semiquinone was also suggested by preliminary EPR data. The results suggest that electron transfer from the iron-sulphur clusters (N2) to Coenzyme Q occurs in two steps gated by two different conformations, the former being sensitive to Rotenone and the latter to Stigmatellin.  相似文献   

2.
According to the 'mitochondrial theory of aging' it is expected that the activity of NADH Coenzyme Q reductase (Complex I) would be most severely affected among mitochondrial enzymes, since mitochondrial DNA encodes for 7 subunits of this enzyme. Being these subunits the site of binding of the acceptor substrate (Coenzyme Q) and of most inhibitors of the enzyme, it is also expected that subtle kinetic changes of quinone affinity and enzyme inhibition could develop in aging before an overall loss of activity would be observed.The overall activity of Complex I was decreased in several tissues from aged rats, nevertheless it was found that direct assay of Complex I using artificial quinone acceptors may underevaluate the enzyme activity. The most acceptable results could be obtained by applying the 'pool equation' to calculate Complex I activity from aerobic NADH oxidation; using this method it was found that the decrease in Complex I activity in mitochondria from old animals was greater than the activity calculated by direct assay of NADH Coenzyme Q reductase.A decrease of NADH oxidation and its rotenone sensitivity was observed in nonsynaptic mitochondria, but not in synaptic 'light' and 'heavy' mitochondria of brain cortex from aged rats.In a study of Complex I activity in human platelet membranes we found that the enzyme activity was unchanged but the titre for half-inhibition by rotenone was significantly increased in aged individuals and proposed this change as a suitable biomarker of aging and age-related diseases. (Mol Cell Biochem 174: 329–333, 1997)  相似文献   

3.
Mitochondrial Complex I (NADH Coenzyme Q oxidoreductase) is the least understood of respiratory complexes. In this review we emphasize some novel findings on this enzyme that are of relevance to the pathogenesis of neurodegenerative diseases. Besides Coenzyme Q (CoQ), also oxygen may be an electron acceptor from the enzyme, with generation of superoxide radical in the mitochondrial matrix. The site of superoxide generation is debated: we present evidence based on the rational use of several inhibitors that the one-electron donor to oxygen is an iron-sulphur cluster, presumably N2. On this assumption we present a novel mechanism of electron transfer to the acceptor, CoQ. Strong evidence is accumulating that electron transfer from Complex I to Complex III via CoQ is not performed by operation of the CoQ pool but by direct channelling within a super-complex including Complex I, Complex III and bound CoQ. Besides structural evidence of a Complex I -Complex III aggregate obtained by native electrophoresis, we have obtained kinetic evidence based on metabolic flux analysis, demonstrating that Complexes I and III behave as an individual enzyme. Quantitative and qualitative changes of phospholipids, including peroxidation, may affect the supercomplex formation. Complex I is deeply involved in pathological changes, including neurodegeneration. Maternally inherited mutations in mitochondrial DNA genes encoding for Complex I subunits are at the basis of Leber's Hereditary Optic Neuropathy; a decrease of electron transfer in the complex, due to the mutations, is not sufficient per se to explain the clinical phenotype, and other factors including proton translocation and oxygen radical generation have been considered of importance. Complex I changes are also involved in more common neurological diseases of the adult and old ages. In this review we discuss Parkinson's disease, where the pathogenic involvement of Complex I is better understood; the accumulated evidence on the mode of action of Complex I inhibitors and their effect on oxygen radical generation is discussed in terms of the aetiology and pathogenesis of the disease.  相似文献   

4.
The quinone binding site (Q-site) of Mitochondrial Complex II (succinate-ubiquinone oxidoreductase) is the target for a number of inhibitors useful for elucidating the mechanism of the enzyme. Some of these have been developed as fungicides or pesticides, and species-specific Q-site inhibitors may be useful against human pathogens. We report structures of chicken Complex II with six different Q-site inhibitors bound, at resolutions 2.0–2.4 Å. These structures show the common interactions between the inhibitors and their binding site. In every case a carbonyl or hydroxyl oxygen of the inhibitor is H-bonded to Tyr58 in subunit SdhD and Trp173 in subunit SdhB. Two of the inhibitors H-bond Ser39 in subunit SdhC directly, while two others do so via a water molecule. There is a distinct cavity that accepts the 2-substituent of the carboxylate ring in flutolanil and related inhibitors. A hydrophobic “tail pocket” opens to receive a side-chain of intermediate-length inhibitors. Shorter inhibitors fit entirely within the main binding cleft, while the long hydrophobic side chains of ferulenol and atpenin A5 protrude out of the cleft into the bulk lipid region, as presumably does that of ubiquinone. Comparison of mitochondrial and Escherichia coli Complex II shows a rotation of the membrane-anchor subunits by 7° relative to the iron?sulfur protein. This rotation alters the geometry of the Q-site and the H-bonding pattern of SdhB:His216 and SdhD:Asp57. This conformational difference, rather than any active-site mutation, may be responsible for the different inhibitor sensitivity of the bacterial enzyme.  相似文献   

5.
Mitochondrial Complex I [NADH Coenzyme Q (CoQ) oxidoreductase] is the least understood of respiratory complexes. In this review we emphasize some novel findings on this enzyme that are of relevance to the pathogenesis of neurodegenerative diseases. Besides CoQ, also oxygen may be an electron acceptor from the enzyme, with generation of superoxide radical in the mitochondrial matrix. The site of superoxide generation is debated: we present evidence based on the rational use of several inhibitors that the one-electron donor to oxygen is an iron-sulphur cluster, presumably N2. On this assumption we present a novel mechanism of electron transfer to the acceptor, CoQ. Complex I is deeply involved in pathological changes, including neurodegeneration. Complex I changes are involved in common neurological diseases of the adult and old ages. Mitochondrial cytopathies due to mutations of either nuclear or mitochondrial DNA may represent a useful model of neurodegeneration. In this review we discuss Parkinson’s disease, where the pathogenic involvement of Complex I is better understood; the accumulated evidence on the mode of action of Complex I inhibitors and their effect on oxygen radical generation is discussed in terms of the aetiology and pathogenesis of the disease. Special issue article in honor of Dr. Anna Maria Giuffrida-Stella.  相似文献   

6.
One of the most serious side effects associated with the therapy of HIV-1 infection is the appearance of viral strains that exhibit resistance to protease inhibitors. The active site mutant V82F/I84V has been shown to lower the binding affinity of protease inhibitors in clinical use. To identify the origin of this effect, we have investigated the binding thermodynamics of the protease inhibitors indinavir, ritonavir, saquinavir, and nelfinavir to the wild-type HIV-1 protease and to the V82F/I84V resistant mutant. The main driving force for the binding of all four inhibitors is a large positive entropy change originating from the burial of a significant hydrophobic surface upon binding. At 25 degrees C, the binding enthalpy is unfavorable for all inhibitors except ritonavir, for which it is slightly favorable (-2.3 kcal/mol). Since the inhibitors are preshaped to the geometry of the binding site, their conformational entropy loss upon binding is small, a property that contributes to their high binding affinity. The V82F/I84V active site mutation lowers the affinity of the inhibitors by making the binding enthalpy more positive and making the entropy change slightly less favorable. The effect on the enthalpy change is, however, the major one. The predominantly enthalpic effect of the V82F/I84V mutation is consistent with the idea that the introduction of the bulkier Phe side chain at position 82 and the Val side chain at position 84 distort the binding site and weaken van der Waals and other favorable interactions with inhibitors preshaped to the wild-type binding site. Another contribution of the V82F/I84V to binding affinity originates from an increase in the energy penalty associated with the conformational change of the protease upon binding. The V82F/I84V mutant is structurally more stable than the wild-type protease by about 1.4 kcal/mol. This effect, however, affects equally the binding affinity of substrate and inhibitors.  相似文献   

7.
Reactive oxygen species (ROS) production by respiratory Complex I from Escherichia coli was studied in bacterial membrane fragments and in the isolated and purified enzyme, either solubilized or incorporated in proteoliposomes. We found that the replacement of a single amino acid residue in close proximity to the nicotinamide adenine dinucleotide (NADH)‐binding catalytic site (E95 in the NuoF subunit) dramatically increases the reactivity of Complex I towards dioxygen (O2). In the E95Q variant short‐chain ubiquinones exhibit strong artificial one‐electron reduction at the catalytic site, also leading to a stronger increase in ROS production. Two mechanisms can contribute to the observed kinetic effects: (a) a change in the reactivity of flavin mononucleotide (FMN) towards dioxygen at the catalytic site, and (b) a change in the population of the ROS‐generating state. We propose the existence of two (closed and open) states of the NAD+‐bound enzyme as one feature of the substrate‐binding site of Complex I. The analysis of the kinetic model of ROS production allowed us to propose that the population of Complex I with reduced FMN is always low in the wild‐type enzyme even at low ambient redox potentials, minimizing the rate of reaction with O2 in contrast to E95Q variant.  相似文献   

8.
Predicting off-targets by computational methods is getting increasing importance in early drug discovery stages. We herewith present a computational method based on binding site three-dimensional comparisons, which prompted us to investigate the cross-reaction of protein kinase inhibitors with synapsin I, an ATP-binding protein regulating neurotransmitter release in the synapse. Systematic pair-wise comparison of the staurosporine-binding site of the proto-oncogene Pim-1 kinase with 6,412 druggable protein-ligand binding sites suggested that the ATP-binding site of synapsin I may recognize the pan-kinase inhibitor staurosporine. Biochemical validation of this hypothesis was realized by competition experiments of staurosporine with ATP-γ35S for binding to synapsin I. Staurosporine, as well as three other inhibitors of protein kinases (cdk2, Pim-1 and casein kinase type 2), effectively bound to synapsin I with nanomolar affinities and promoted synapsin-induced F-actin bundling. The selective Pim-1 kinase inhibitor quercetagetin was shown to be the most potent synapsin I binder (IC50  = 0.15 µM), in agreement with the predicted binding site similarities between synapsin I and various protein kinases. Other protein kinase inhibitors (protein kinase A and chk1 inhibitor), kinase inhibitors (diacylglycerolkinase inhibitor) and various other ATP-competitors (DNA topoisomerase II and HSP-90α inhibitors) did not bind to synapsin I, as predicted from a lower similarity of their respective ATP-binding sites to that of synapsin I. The present data suggest that the observed downregulation of neurotransmitter release by some but not all protein kinase inhibitors may also be contributed by a direct binding to synapsin I and phosphorylation-independent perturbation of synapsin I function. More generally, the data also demonstrate that cross-reactivity with various targets may be detected by systematic pair-wise similarity measurement of ligand-annotated binding sites.  相似文献   

9.
Complex I (NADH:ubiquinone oxidoreductase) has a central function in oxidative phosphorylation and hence for efficient ATP production in most prokaryotic and eukaryotic cells. This huge membrane protein complex transfers electrons from NADH to ubiquinone and couples this exergonic redox reaction to endergonic proton pumping across bioenergetic membranes. Although quinone reduction seems to be critical for energy conversion, this part of the reaction is least understood. Here we summarize and discuss experimental evidence indicating that complex I contains an extended ubiquinone binding pocket at the interface of the 49-kDa and PSST subunits. Close to iron–sulfur cluster N2, the proposed immediate electron donor for ubiquinone, a highly conserved tyrosine constitutes a critical element of the quinone reduction site. A possible quinone exchange path leads from cluster N2 to the N-terminal β-sheet of the 49-kDa subunit. We discuss the possible functions of a highly conserved HRGXE motif and a redox–Bohr group associated with cluster N2. Resistance patterns observed with a large number of point mutations suggest that all types of hydrophobic complex I inhibitors also act at the interface of the 49-kDa and the PSST subunit. Finally, current controversies regarding the number of ubiquinone binding sites and the position of the site of ubiquinone reduction are discussed.  相似文献   

10.
The synthetic retinoid N-(4-hydroxyphenyl)retinamide (4HPR) is being examined in both chemoprevention and therapy clinical trials. Yet, its mechanism(s) of action is still not fully elucidated. In previous studies, an increase in mitochondrial reactive oxygen species has been proposed as one mechanism through which 4HPR could exert its proapoptotic effects. This study explored whether mitochondrial respiration is required for 4HPR action using human cutaneous squamous cell carcinoma cells and respiration-deficient clones. In parental cells, 4HPR rapidly promoted hydroperoxide production followed by mitochondrial permeability transition, caspase activity, and DNA fragmentation. Short term exposure to 4HPR also inhibited oxygen consumption in parental cells. This activity was reversed by the antioxidant vitamin C indicating the prooxidant effect of 4HPR directly impaired mitochondrial function. In respiration-deficient clones, the proapoptotic qualities of 4HPR were conspicuously diminished illustrating a central role for mitochondrial respiration in 4HPR-induced cell death. In parental cells, various mitochondrial inhibitors were examined to determine potential sites associated with the prooxidant activity of 4HPR. Inhibitors of Complex II as well as center i inhibitors of Complex III enhanced 4HPR-induced hydroperoxide production. Complex I inhibitors, center o inhibitors of Complex III, cyanide, oligomycin A, and coenzyme Q analogues decreased 4HPR-induced hydroperoxide production. The coenzyme Q analogues were very effective in this respect, and they also blocked the enhanced hydroperoxide production obtained when center i inhibitors were combined with 4HPR. These results suggest the prooxidant property of 4HPR is associated with redox metabolism via an enzymatic process occurring at a quinone-binding site in Complex I and/or center o of Complex III.  相似文献   

11.
The addition of NADH to submitochondrial particles inhibited by agents which interrupt electron transport from NADH-Q oxidoreductase (Complex I) to Q10 (rotenone, piericidin A, and MPP+) results in superoxide formation and lipid peroxidation. A study of the quantitative relations now shows that oxyradical formation does not appear to be the direct result of the inhibition. Although tetraphenyl boron (TPB) greatly enhances the inhibition by MPP+, it has no effect on O2. formation or lipid peroxidation. When submitochondrial particles completely inhibited by rotenone or piericidin A are treated with bovine serum albumin to remove spuriously bound inhibitor molecules without affecting those bound at the specific inhibition site, NADH-Q activity remains inhibited and lipid peroxidation occurs but superoxide formation ceases. Thus oxyradical formation may be the result of the binding of inhibitors at sites in the membrane other than those related to the inhibition of electron transport.  相似文献   

12.
Iacob RE  Zhang J  Gray NS  Engen JR 《PloS one》2011,6(1):e15929
Abl kinase inhibitors targeting the ATP binding pocket are currently employed as potent anti-leukemogenic agents but drug resistance has become a significant clinical limitation. Recently, a compound that binds to the myristate pocket of Abl (GNF-5) was shown to act cooperatively with nilotinib, an ATP-competitive inhibitor to target the recalcitrant “T315I” gatekeeper mutant of Bcr-Abl. To uncover an explanation for how drug binding at a distance from the kinase active site could lead to inhibition and how inhibitors could combine their effects, hydrogen exchange mass spectrometry (HX MS) was employed to monitor conformational effects in the presence of both dasatinib, a clinically approved ATP-site inhibitor, and GNF-5. While dasatinib binding to wild type Abl clearly influenced Abl conformation, no binding was detected between dasatinib and T315I. GNF-5, however, elicited the same conformational changes in both wild type and T315I, including changes to dynamics within the ATP site located approximately 25 Å from the site of GNF-5 interaction. Simultaneous binding of dasatinib and GNF-5 to T315I caused conformational and/or dynamics changes in Abl such that effects of dasatinib on T315I were the same as when it bound to wild type Abl. These results provide strong biophysical evidence that allosteric interactions play a role in Abl kinase downregulation and that targeting sites outside the ATP binding site can provide an important pharmacological tool to overcome mutations that cause resistance to ATP-competitive inhibitors.  相似文献   

13.
Complex I in bovine heart submitochondrial particles catalyses the NADH-supported generation of superoxide anion; adrenaline is oxidised by superoxide to adrenochrome that, on its hand, is reduced by Complex I, thus establishing a redox cycle that amplifies the superoxide production. The routes in Complex I for superoxide formation and for adrenochrome reduction appear to be different, since they have a different sensitivity to Complex I inhibitors. The results are discussed in terms of current assays for superoxide detection and of pathologies linked to catecholamine oxidation.  相似文献   

14.
The mitochondrial respiratory chain is a powerful source of reactive oxygen species, considered as the pathogenic agent of many diseases and of aging. We have investigated the role of Complex I in superoxide radical production in bovine heart submitochondrial particles and found, by combined use of specific inhibitors of Complex I and by Coenzyme Q (CoQ) extraction from the particles, that the one-electron donor in the Complex to oxygen is a redox center located prior to the binding sites of three different types of CoQ antagonists, to be identified with a Fe-S cluster, most probably N2 on the basis of several known properties of this cluster. Short chain CoQ analogs enhance superoxide formation, presumably by mediating electron transfer from N2 to oxygen. The clinically used CoQ analog, idebenone, is particularly effective in promoting superoxide formation.  相似文献   

15.
This review examines two aspects of the structure and function of mitochondrial Complex I (NADH Coenzyme Q oxidoreductase) that have become matter of recent debate. The supramolecular organization of Complex I and its structural relation with the remainder of the respiratory chain are uncertain. Although the random diffusion model [C.R. Hackenbrock, B. Chazotte, S.S. Gupte, The random collision model and a critical assessment of diffusion and collision in mitochondrial electron transport, J. Bioenerg. Biomembranes 18 (1986) 331-368] has been widely accepted, recent evidence suggests the presence of supramolecular aggregates. In particular, evidence for a Complex I-Complex III supercomplex stems from both structural and kinetic studies. Electron transfer in the supercomplex may occur by electron channelling through bound Coenzyme Q in equilibrium with the pool in the membrane lipids. The amount and nature of the lipids modify the aggregation state and there is evidence that lipid peroxidation induces supercomplex disaggregation. Another important aspect in Complex I is its capacity to reduce oxygen with formation of superoxide anion. The site of escape of the single electron is debated and either FMN, iron-sulphur clusters, and ubisemiquinone have been suggested. The finding in our laboratory that two classes of hydrophobic inhibitors have opposite effects on superoxide production favours an iron-sulphur cluster (presumably N2) is the direct oxygen reductant. The implications in human pathology of better knowledge on these aspects of Complex I structure and function are briefly discussed.  相似文献   

16.
Muzammil S  Ross P  Freire E 《Biochemistry》2003,42(3):631-638
A major problem in the chemotherapy of HIV-1 infection is the appearance of drug resistance. In the case of HIV-1 protease inhibitors, resistance originates from mutations in the protease molecule that lower the affinity of inhibitors while still maintaining a viable enzymatic profile. Drug resistance mutations can be classified as active site or non-active site mutations depending on their location within the protease molecule. Active site mutations directly affect drug/target interactions, and their action can be readily understood in structural terms. Non-active site mutations influence binding from distal locations, and their mechanism of action is not immediately apparent. In this paper, we have characterized a mutant form of the HIV-1 protease, ANAM-11, identified in clinical isolates from HIV-1 infected patients treated with protease inhibitors. This mutant protease contains 11 mutations, 10 of which are located outside the active site (L10I/M36I/S37D/M46I/R57K/L63P/A71V/G73S/L90M/I93L) and 1 within the active site (I84V). ANAM-11 lowers the binding affinity of indinavir, nelfinavir, saquinavir, and ritonavir by factors of 4000, 3300, 5800, and 80000, respectively. Surprisingly, most of the loss in inhibitor affinity is due to the non-active site mutations as demonstrated by additional experiments performed with a protease containing only the 10 non-active site mutations (NAM-10) and another containing only the active site mutation (A-1). Kinetic analysis with two different substrates yielded comparable catalytic efficiencies for A-1, ANAM-11, NAM-10, and the wild-type protease. These studies demonstrate that non-active site mutations can be the primary source of resistance and that their role is not necessarily limited to compensate deleterious effects of active site mutations. Analysis of the structural stability of the proteases by differential scanning calorimetry reveals that ANAM-11 and NAM-10 are structurally more stable than the wild-type protease while A-1 is less stable. Together, the binding and structural thermodynamic results suggest that the non-active site mutants affect inhibitor binding by altering the geometry of the binding site cavity through the accumulation of mutations within the core of the protease molecule.  相似文献   

17.
The Tn3-encoded resolvase protein promotes a site-specific recombination reaction between two directly repeated copies of the recombination site res. Several inhibitors that block this event in vitro have been isolated. In this study four of these inhibitors were tested on various steps in the recombination reaction. Two inhibitors. A9387 and A1062, inhibit resolvase binding to the res site. Further, DNase I footprinting revealed that at certain concentrations of A9387 and A1062, resolvase was preferentially bound to site I of res, the site containing the recombinational crossover point. The two other inhibitors, A20812 and A21960, do not affect resolvase binding and bending of the DNA but inhibit synapse formation between resolvase and two directly repeated res sites.  相似文献   

18.
The majority of kinase inhibitors that have been developed so far--known as type I inhibitors--target the ATP binding site of the kinase in its active conformation, in which the activation loop is phosphorylated. Recently, crystal structures of inhibitors such as imatinib (STI571), BIRB796 and sorafenib (BAY43-9006)--known as type II inhibitors--have revealed a new binding mode that exploits an additional binding site immediately adjacent to the region occupied by ATP. This pocket is made accessible by an activation-loop rearrangement that is characteristic of kinases in an inactive conformation. Here, we present a structural analysis of binding modes of known human type II inhibitors and demonstrate that they conform to a pharmacophore model that is currently being used to design a new generation of kinase inhibitors.  相似文献   

19.
Studies of the structure–activity relationships of ubiquinones and specific inhibitors are helpful to probe the structural and functional features of the ubiquinone reduction site of bovine heart mitochondrial complex I. Bulky exogenous short-chain ubiquinones serve as sufficient electron acceptors from the physiological ubiquinone reduction site of bovine complex I. This feature is in marked contrast to other respiratory enzymes such as mitochondrial complexes II and III. For various complex I inhibitors, including the most potent inhibitors, acetogenins, the essential structural factors that markedly affect the inhibitory potency are not necessarily obvious. Thus, the loose recognition by the enzyme of substrate and inhibitor structures may reflect the large cavitylike structure of the ubiquinone (or inhibitor) binding domain in the enzyme. On the other hand, several phenomena are difficult to explain by a simple one-catalytic site model for ubiquinone.  相似文献   

20.
The replication initiator protein of bacteriophage f1 (gene II protein) binds to the phage origin and forms two complexes that are separable by polyacrylamide gel electrophoresis. Complex I is formed at low gene II protein concentrations, and shows protection from DNase I of about 25 base-pairs (from position +2 to +28 relative to the nicking site) at the center of the minimal origin sequence. Complex II is produced at higher concentrations of the protein, and has about 40 base-pairs (from -7 to +33) protected. On the basis of gel mobility, complex II appears to contain twice the amount of gene II protein as does complex I. The 40 base-pair sequence protected in complex II corresponds to the minimal origin sequence as determined by in-vivo analyses. The central 15 base-pair sequence (from +6 to +20) of the minimal origin consists of two repeats in inverted orientation. This sequence, when cloned into a plasmid, can form complex I, but not complex II. We call this 15 base-pair element the core binding sequence for gene II protein. Methylation interference with the formation of complex I by the wild-type origin indicates that gene II protein contacts six guanine residues located in a symmetric configuration within the core binding sequence. Formation of complex II requires, in addition to the core binding sequence, the adjacent ten base-pair sequence on the right containing a third homologous repeat. A methylation interference experiment performed on complex II indicates that gene II protein interacts homologously with the three repeats. In complex II, gene II protein protects from DNase I digestion not only ten base-pairs on the right but also ten base-pairs on the left of the sequence that is protected in complex I. Footprint analyses of various deletion mutants indicate that the left-most ten base-pairs are protected regardless of their sequence. The site of nicking by gene II protein is located within this region. A model is presented for the binding reaction involving both protein-DNA and protein-protein interactions.  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