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 642 毫秒
1.
The dynamic interactions between leukocyte integrin receptors and ligands in the vascular endothelium, extracellular matrix, or invading pathogens result in leukocyte adhesion, extravasation, and phagocytosis. This work examined the mechanical strength of the connection between iC3b, a complement component that stimulates phagocytosis, and the ligand‐binding domain, the I‐domain, of integrin αMβ2. Single‐molecule force measurements of αM I‐domain–iC3b complexes were conducted by atomic force microscope. Strikingly, depending on loading rates, immobilization of the I‐domain via its C‐terminus resulted in a 1.3‐fold to 1.5‐fold increase in unbinding force compared with I‐domains immobilized via the N‐terminus. The force spectra (unbinding force versus loading rate) of the I‐domain–iC3b complexes revealed that the enhanced mechanical strength is due to a 2.4‐fold increase in the lifetime of the I‐domain–iC3b bond. Given the structural and functional similarity of all integrin I‐domains, our result supports the existing allosteric regulatory model by which the ligand binding strength of integrin can be increased rapidly when a force is allowed to stretch the C‐terminus of the I‐domain. This type of mechanism may account for the rapid ligand affinity adjustment during leukocyte migration. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

2.
Molecular recognition force spectroscopy (MR‐FS) was applied to investigate the dynamic interaction between aptamer GBI‐10 and tenascin‐C (TN‐C) on human glioblastoma cell surface at single‐molecule level. The unbinding force between aptamer GBI‐10 and TN‐C was 39 pN at the loading rate of 0.3 nN sec?1. A series of kinetic parameters concerning interaction process such as the unbinding force fu, the association rate constant kon, dissociation rate constant at zero force koff, and dissociation constant KD for aptamer GBI‐10/TN‐C complexes were acquired. In addition, the interaction of aptamer GBI‐10 with TN‐C depended on the presence of Mg2+. This work demonstrates that MR‐FS can be used as an attractive tool for exploring the interaction forces and dynamic process of aptamer and ligand at the single‐molecule level. As a future perspective, MR‐FS may be used as a potential diagnostic and therapeutic tool by combining with other techniques. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

3.
We study the unbiased folding/unfolding thermodynamics of the Trp‐cage miniprotein using detailed molecular dynamics simulations of an all‐atom model of the protein in explicit solvent using the Amberff99SB force field. Replica‐exchange molecular dynamics simulations are used to sample the protein ensembles over a broad range of temperatures covering the folded and unfolded states at two densities. The obtained ensembles are shown to reach equilibrium in the 1 μs/replica timescale. The total simulation time used in the calculations exceeds 100 μs. Ensemble averages of the fraction folded, pressure, and energy differences between the folded and unfolded states as a function of temperature are used to model the free energy of the folding transition, ΔG(P, T), over the whole region of temperatures and pressures sampled in the simulations. The ΔG(P, T) diagram describes an ellipse over the range of temperatures and pressures sampled, predicting that the system can undergo pressure‐induced unfolding and cold denaturation at low temperatures and high pressures, and unfolding at low pressures and high temperatures. The calculated free energy function exhibits remarkably good agreement with the experimental folding transition temperature (Tf = 321 K), free energy, and specific heat changes. However, changes in enthalpy and entropy are significantly different than the experimental values. We speculate that these differences may be due to the simplicity of the semiempirical force field used in the simulations and that more elaborate force fields may be required to describe appropriately the thermodynamics of proteins. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
The autotrophic growth of an oil‐rich indigenous microalgal isolate, identified as Chlorella vulgaris C? C, was promoted by using engineering strategies to obtain the microalgal oil for biodiesel synthesis. Illumination with a light/dark cycle of 14/10 (i.e., 14 h light‐on and 10 h light‐off) resulted in a high overall oil production rate (voil) of 9.78 mg/L/day and a high electricity conversion efficiency (Ec) of 23.7 mg cell/kw h. When using a NaHCO3 concentration of 1,500 mg/L as carbon source, the voil and Ec were maximal at 100 mg/L/day and 128 mg/kw h, respectively. A Monod type model was used to describe the microalgal growth kinetics with an estimated maximum specific growth rate (μmax) of 0.605 day?1 and a half saturation coefficient (Ks) of 124.9 mg/L. An optimal nitrogen source (KNO3) concentration of 625 mg/L could further enhance the microalgal biomass and oil production, leading to a nearly 6.19 fold increase in voil value. © 2010 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

5.
The racemic and enantioselective synthesis of a novel glyceric acid derivative, namely, 2,3‐dihydroxy‐3‐(3,4‐dihydroxyphenyl)‐propionic acid as well as the antioxidant activities is described. The virtually pure enantiomers, (+)‐(2R,3S)‐2,3‐dihydroxy‐3‐(3,4‐dihydroxyphenyl)‐propionic acid and (?)‐(2S,3R)‐2,3‐dihydroxy‐3‐(3,4‐dihydroxyphenyl)‐propionic acid were synthesized for the first time via Sharpless asymmetric dihydroxylation of trans‐caffeic acid derivatives using the enantiocomplementary catalysts, (DHQD)2‐PHAL and (DHQ)2‐PHAL. The determination of enantiomeric purity of the novel chiral glyceric acid derivatives was performed by high‐performance liquid chromatographic techniques on the stage of their alkylated precursors. The novel glyceric acid derivatives show strong antioxidant activity against hypochlorite and N,N‐diphenyl‐N‐picryl‐hydrazyl free radical. Their antioxidant activity is about 40‐fold higher than that of the corresponding natural polyether and three‐fold higher of trans‐caffeic acid itself. Chirality, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

6.
Helices are important structural/recognition elements in proteins and peptides. Stability and conformational differences between helices composed of α‐ and β‐amino acids as scaffolds for mimicry of helix recognition has become a theme in medicinal chemistry. Furthermore, helices formed by β‐amino acids are experimentally more stable than those formed by α‐amino acids. This is paradoxical because the larger sizes of the hydrogen‐bonding rings required by the extra methylene groups should lead to entropic destabilization. In this study, molecular dynamics simulations using the second‐generation force field, AMOEBA (Ponder, J.W., et al., Current status of the AMOEBA polarizable force field. J Phys Chem B, 2010. 114 (8): p. 2549–64.) explored the stability and hydrogen‐bonding patterns of capped oligo‐β‐alanine, oligoalanine, and oligoglycine dodecamers in water. The MD simulations showed that oligo‐β‐alanine has strong acceptor+2 hydrogen bonds, but surprisingly did not contain a large content of 312‐helical structures, possibly due to the sparse distribution of the 312‐helical structure and other structures with acceptor+2 hydrogen bonds. On the other hand, despite its backbone flexibility, the β‐alanine dodecamer had more stable and persistent <3.0 Å hydrogen bonds. Its structure was dominated more by multicentered hydrogen bonds than either oligoglycine or oligoalanine helices. The 31 (PII) helical structure, prevalent in oligoglycine and oligoalanine, does not appear to be stable in oligo‐β‐alanine indicating its competition with other structures (stacking structure as indicated by MD analyses). These differences are among the factors that shape helical structural preferences and the relative stabilities of these three oligopeptides. Proteins 2014; 82:3043–3061. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
In aqueous solutions of the amylase–iodine complex the concentration of free iodine [If]v after reaching equilibrium (or closely approximating it) is determined by the following factors: temperature, pH, concentration of iodide ions and amylose, and DP of amylose. In the present paper the role of temperature, amylose concentration, and DP has been investigated. At half-saturation of amylose by iodine, the reciprocal value of free iodine defines the equilibrium constant: 1/[If]v = K. The relation between [If]v, in normality and temperature is the following: 5 + log [If]v = ?(2.132/T) + 8.52, for DP n = 1290, 0.4 mg. amylose in 100 ml. 0.1N HCl. The value of the energy of activation Ea between 2 and 52°C. is 9.72 kcal./mole. The influence of amylose concentration [Am] on photometrically determined [If]v, at 20°C, in the range of 0.1–1.2 mg./100 ml. 0.1 N HCl for DP n = 1290 is: 5 + log [If]v = 0.209 ? 0.047 log [Am]. At [Am] = 0.6 mg. amylose/ 100 ml. 0.1 N HCl and 20°C, the value of [If]v depends on DP n as follows: 5 + log [If]v = 0.085 = + 0.222 log (104/DP n). These above equations are summarized by the relation: [If]v = exp {16.865 ? (Ea/RT)}[Am]0.047(104/DP n)0.222 ×10?5 Considering that the determination of [If]v by automatic photometric titration can be performed quickly and with appropriate reproducibility, this method is convenient for a rapid empirical and approximate determination of DP of amylose on a microscale. The iodine-binding capacity [IBC] as well as the value of λmax, have been also investigated as functions of DP n, by photometric and by amperometric titration.  相似文献   

8.
The tyrosinase gene from Ralstonia solanacearum (GenBank NP518458) was subjected to random mutagenesis resulting in tyrosinase variants (RVC10 and RV145) with up to 3.2‐fold improvement in kcat, 5.2‐fold lower Km and 16‐fold improvement in catalytic efficiency for D ‐tyrosine. Based on RVC10 and RV145 mutated sequences, single mutation variants were generated with all variants showing increased kcat for D ‐tyrosine compared to the wild type (WT). All single mutation variants based on RV145 had a higher kcat and Km value compared to the RV145 and thus the combination of four mutations in RV145 was antagonistic for turnover, but synergistic for affinity of the enzyme for D ‐tyrosine. Single mutation variant 145_V153A exhibited the highest (6.9‐fold) improvement in kcat and a 2.4‐fold increase in Km compared to the WT. Two single mutation variants, C10_N322S and C10_T183I reduced the Km up to 2.6‐fold for D ‐tyrosine but one variant 145_V153A increased the Km 2.4‐fold compared to the WT. Homology based modeling of R. solanacearum tyrosinase showed that mutation V153A disrupts the van der Waals interactions with an α‐helix providing one of the conserved histidine residues of the active site. The kcat and Km values for L ‐tyrosine decreased for RV145 and RVC10 compared to the WT. RV145 exhibited a 2.1‐fold high catalytic efficiency compared to the WT which is a 7.6‐fold lower improvement compared to D ‐tyrosine. RV145 exhibited a threefold higher monophenolase:diphenolase activity ratio for D ‐tyrosine:D ‐DOPA and a 1.4‐fold higher L ‐tyrosine:L ‐DOPA activity ratio compared to the WT. Biotechnol. Bioeng. 2013; 110: 1849–1857. © 2013 Wiley Periodicals, Inc.  相似文献   

9.
Zhang Y  Sun G  Lü S  Li N  Long M 《Biophysical journal》2008,95(11):5439-5448
Forced dissociation of selectin-ligand bonds is crucial to such biological processes as leukocyte recruitment, thrombosis formation, and tumor metastasis. Although the bond rupture has been well known at high loading rate rf (≥102 pN/s), defined as the product of spring constant k and retract velocity v, how the low rf (<102 pN/s) or the low k regulates the bond dissociation remains unclear. Here an optical trap assay was used to quantify the bond rupture at rf ≤ 20 pN/s with low k (∼10−3-10−2 pN/nm) when P-selectin and P-selectin glycoprotein ligand 1 (PSGL-1) were respectively coupled onto two glass microbeads. Our data indicated that the bond rupture force f retained the similar values when rf increased up to 20 pN/s. It was also found that f varied with different combinations of k and v even at the same rf. The most probable force, f*, was enhanced with the spring constant when k < 47.0 × 10−3 pN/nm, indicating that the bond dissociation at low rf was spring constant dependent and that bond rupture force depended on both the loading rate and the mechanical compliance of force transducer. These results provide new insights into understanding the P-selectin glycoprotein ligand 1 bond dissociation at low rf or k.  相似文献   

10.
The fungus Candida albicans is the most common cause of mycotic infections in immunocompromised hosts. Little is known about the initial interactions between Candida and immune cell receptors, such as the C‐type lectin dendritic cell‐specific intracellular cell adhesion molecule‐3 (ICAM‐3)‐grabbing non‐integrin (DC‐SIGN), because a detailed characterization at the structural level is lacking. DC‐SIGN recognizes specific Candida‐associated molecular patterns, that is, mannan structures present in the cell wall of Candida. The molecular recognition mechanism is however poorly understood. We postulated that small differences in mannan‐branching may result in considerable differences in the binding affinity. Here, we exploit atomic force microscope‐based dynamic force spectroscopy with single Candida cells to gain better insight in the carbohydrate recognition capacity of DC‐SIGN. We demonstrate that slight differences in the N‐mannan structure of Candida, that is, the absence or presence of a phosphomannan side chain, results in differences in the recognition by DC‐SIGN as follows: (i) it contributes to the compliance of the outer cell wall of Candida, and (ii) its presence results in a higher binding energy of 1.6 kBT. The single‐bond affinity of tetrameric DC‐SIGN for wild‐type C. albicans is ~10.7 kBT and a dissociation constant kD of 23 μM, which is relatively strong compared with other carbohydrate–protein interactions described in the literature. In conclusion, this study shows that DC‐SIGN specifically recognizes mannan patterns on C. albicans with high affinity. Knowledge on the binding pocket of DC‐SIGN and its pathogenic ligands will lead to a better understanding of how fungal‐associated carbohydrate structures are recognized by receptors of the immune system and can ultimately contribute to the development of new anti‐fungal drugs. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

11.
Here, we study mechanical properties of eight 3‐helix proteins (four right‐handed and four left‐handed ones), which are similar in size under stretching at a constant speed and at a constant force on the atomic level using molecular dynamics simulations. The analysis of 256 trajectories from molecular dynamics simulations with explicit water showed that the right‐handed three‐helix domains are more mechanically resistant than the left‐handed domains. Such results are observed at different extension velocities studied (192 trajectories obtained at the following conditions: v = 0.1, 0.05, and 0.01 Å ps?1, T = 300 K) and under constant stretching force (64 trajectories, F = 800 pN, T = 300 K). We can explain this by the fact, at least in part, that the right‐handed domains have a larger number of contacts per residue and the radius of cross section than the left‐handed domains. Proteins 2014; 82:90–102. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
The global fold of human cannabinoid type 2 (CB2) receptor in the agonist‐bound active state in lipid bilayers was investigated by solid‐state 13C‐ and 15N magic‐angle spinning (MAS) NMR, in combination with chemical‐shift prediction from a structural model of the receptor obtained by microsecond‐long molecular dynamics (MD) simulations. Uniformly 13C‐ and 15N‐labeled CB2 receptor was expressed in milligram quantities by bacterial fermentation, purified, and functionally reconstituted into liposomes. 13C MAS NMR spectra were recorded without sensitivity enhancement for direct comparison of Cα, Cβ, and C?O bands of superimposed resonances with predictions from protein structures generated by MD. The experimental NMR spectra matched the calculated spectra reasonably well indicating agreement of the global fold of the protein between experiment and simulations. In particular, the 13C chemical shift distribution of Cα resonances was shown to be very sensitive to both the primary amino acid sequence and the secondary structure of CB2. Thus the shape of the Cα band can be used as an indicator of CB2 global fold. The prediction from MD simulations indicated that upon receptor activation a rather limited number of amino acid residues, mainly located in the extracellular Loop 2 and the second half of intracellular Loop 3, change their chemical shifts significantly (≥1.5 ppm for carbons and ≥5.0 ppm for nitrogens). Simulated two‐dimensional 13Cα(i)? 13C?O(i) and 13C?O(i)? 15NH(i + 1) dipolar‐interaction correlation spectra provide guidance for selective amino acid labeling and signal assignment schemes to study the molecular mechanism of activation of CB2 by solid‐state MAS NMR. Proteins 2014; 82:452–465. © 2013 Wiley Periodicals, Inc.  相似文献   

13.
Indoxacarb and metaflumizone belong to a relatively new class of sodium channel blocker insecticides (SCBIs). Due to intensive use of indoxacarb, field‐evolved indoxacarb resistance has been reported in several lepidopteran pests, including the diamondback moth Plutella xylostella, a serious pest of cruciferous crops. In particular, the BY12 population of P. xylostella, collected from Baiyun, Guangdong Province of China in 2012, was 750‐fold more resistant to indoxacarb and 70‐fold more resistant to metaflumizone compared with the susceptible Roth strain. Comparison of complementary DNA sequences encoding the sodium channel genes of Roth and BY12 revealed two point mutations (F1845Y and V1848I) in the sixth segment of domain IV of the PxNav protein in the BY population. Both mutations are located within a highly conserved sequence region that is predicted to be involved in the binding sites of local anesthetics and SCBIs based on mammalian sodium channels. A significant correlation was observed among 10 field‐collected populations between the mutant allele (Y1845 or I1848) frequencies (1.7% to 52.5%) and resistance levels to both indoxacarb (34‐ to 870‐fold) and metaflumizone (1‐ to 70‐fold). The two mutations were never found to co‐exist in the same allele of PxNav, suggesting that they arose independently. This is the first time that sodium channel mutations have been associated with high levels of resistance to SCBIs. F1845Y and V1848I are molecular markers for resistance monitoring in the diamondback moth and possibly other insect pest species.  相似文献   

14.
In circulation, cancer cells induce platelet activation, leading to the formation of a cancer cell‐encircling platelet cloak which facilitates each step of the metastatic cascade. Since cancer patients treated with the anticoagulant heparin showed reduced metastasis rates and improved survival, it is supposed that heparin suppresses the cloak's formation by inhibiting the interaction between platelet's adhesion molecule P‐selectin with its ligands on cancer cells. To quantify this heparin effect, we developed a single‐cell force spectroscopy approach and quantified the adhesion (maximum adhesion force [FA] and detachment work [WD]) between platelets and human non‐small cell lung cancer cells (A549). A configuration was used in which A549 cells were glued to tipless cantilevers and force‐distance (F‐D) curves were recorded on a layer of activated platelets. The concentration‐response relationship was determined for heparin at concentrations between 1 and 100 U/mL. Sigmoid dose‐response fit revealed half‐maximal inhibitory concentration (IC50) values of 8.01 U/mL (FA) and 6.46 U/mL (WD) and a maximum decrease of the adhesion by 37.5% (FA) and 38.42% (WD). The effect of heparin on P‐selectin was tested using anti‐P‐selectin antibodies alone and in combination with heparin. Adding heparin after antibody treatment resulted in an additional reduction of 9.52% (FA) and 7.12% (WD). Together, we quantified heparin's antimetastatic effect and proved that it predominantly is related to the blockage of P‐selectin. Our approach represents a valuable method to investigate the adhesion of platelets to cancer cells and the efficiency of substances to block this interaction.  相似文献   

15.
Dawei Zhang 《Proteins》2014,82(9):2209-2219
Molecular dynamics (MD) simulation is commonly used in the study of protein dynamics, and in recent years, the extension of MD simulation to the study of metalloproteins is gaining much interest. Choice of force field is crucial in MD studies, and the inclusion of metal centers complicates the process of accurately describing the electrostatic environment that surrounds the redox centre. Herein, we would like to explore the importance of including electrostatic contribution from both protein and solvent in the study of metalloproteins. MD simulations with the implementation of thermodynamic integration will be conducted to model the reduction process of azurin from Pseudomonas aeruginosa. Three charge schemes will be used to derive the partial charges of azurin. These charge schemes differ in terms of the amount of immediate environment, respective to copper, considered during charge fitting, which ranges from the inclusion of copper and residues in the first coordination sphere during density functional theory charge fitting to the comprehensive inclusion of protein and solvent effect surrounding the metal centre using polarized protein‐specific charge scheme. From the simulations conducted, the relative reduction potential of the mutated azurins respective to that of wild‐type azurin (ΔEcal) were calculated and compared with experimental values. The ΔEcal approached experimental value with increasing consideration of environmental effect hence substantiating the importance of polarization effect in the study of metalloproteins. This study also attests the practicality of polarized protein‐specific charge as a computational tool capable of incorporating both protein environment and solvent effect into MD simulations. Proteins 2014; 82:2209–2219. © 2014 Wiley Periodicals, Inc.  相似文献   

16.
We report the largest eukaryotic genome to date in the monocot Paris japonica (Melanthiaceae, 1C = 152.23 pg), measured using flow cytometry. This value is 15% larger than any previous estimate and extends the range of genome sizes to c. 2400‐fold across angiosperms and c. 66 000‐fold across eukaryotes. © 2010 The Linnean Society of London, Botanical Journal of the Linnean Society, 2010, 164 , 10–15.  相似文献   

17.
Five ruthenium(II) complexes, i.e., [Ru(bpy)2(TIP)]2+ (bpy=2,2′‐bipyridine; TIP=2‐thiophenimidazo[4,5‐f] [1,10]phenanthroline; 1 ), [Ru(bpy)2(5‐NTIP)]2+ (5‐NTIP=2‐(5‐nitrothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 2 ), [Ru(bpy)2(5‐MOTIP)]2+ (5‐MOTIP=2‐(5‐methoxythiophen)imidazo[4,5‐f] [1,10]phenanthroline; 3 ), [Ru(bpy)2(5‐BTIP)]2+ (5‐BTIP=2‐(5‐bromothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 4 ), and [Ru(bpy)2(4‐BTIP)]2+ (4‐BTIP=2‐(4‐bromothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 5 ), were synthesized and characterized by elemental analysis and UV/VIS, IR, and 1H‐NMR spectroscopic methods. The photophysical and DNA‐binding properties were investigated by means of UV and fluorescence spectroscopic methods and viscosity measurements, respectively. The results suggest that all five complexes can bind to CT‐DNA with various binding strength. Complexes 2 and 3 showed the strongest and the weakest binding affinity, respectively, among these five complexes. Due to the substituent position of the Br‐atom in the ligand, complex 5 interacted stronger with CT‐DNA than complex 4 . The binding affinities of the complexes decreased in the order 2, 5, 4, 1 , and 3 .  相似文献   

18.
Best RB  Mittal J 《Proteins》2011,79(4):1318-1328
Although it is now possible to fold peptides and miniproteins in molecular dynamics simulations, it is well appreciated that force fields are not all transferable to different proteins. Here, we investigate the influence of the protein force field and the solvent model on the folding energy landscape of a prototypical two‐state folder, the GB1 hairpin. We use extensive replica‐exchange molecular dynamics simulations to characterize the free‐energy surface as a function of temperature. Most of these force fields appear similar at a global level, giving a fraction folded at 300 K between 0.2 and 0.8 in all cases, which is a difference in stability of 2.8 kT, and are generally consistent with experimental data at this temperature. The most significant differences appear in the unfolded state, where there are different residual secondary structures which are populated, and the overall dimensions of the unfolded states, which in most of the force fields are too collapsed relative to experimental Förster Resonance Energy Transfer (FRET) data. Proteins 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

19.
The chronological change of photosynthetic efficiency in a frozen storage treatment of the Japanese Nori cultivation industry was examined in the cultivated red alga, Pyropia yezoensis f. narawaensis (Saga‐#5 Strain, Bangiales) by using pulse‐amplitude fluorometry. During the desiccation process that was conducted after the nursery cultivation season in November, the maximum quantum yield (F v/F m) of the gametophytic sporelings growing on the Nori‐net decreased monotonically with decreasing absolute water content (AWC), and was around 0.1 at 20% AWC. During frozen storage of the Nori‐net, the F v/F m of the frozen gametophyte was low but stable, and ranged between 0.10 ± 0.02 SD and 0.14 ± 0.05 SD. The magnitude of F v/F m for the gametophyte of the freezing treatment, after 10 min and 3 h of immersion in seawater, recovered quickly. After 10 min and 3 h of immersion, these values were 0.29 ± 0.12 SD and 0.47 ± 0.05 SD during the 14 days of freezing treatment, and 0.15 ± 0.02 SD and 0.29 ± 0.04 SD after 71 days of freezing treatment, and suggest that the ability to recover gradually decreased as the storage duration increased. The response of F v/F m from general cultivation (i.e., directly cultivated from the nursery cultivation season) and those after 47 days of freezing were almost identical, suggesting that the current Nori net frozen storage period (6 or 7 weeks) was not detrimental to the gametophyte.  相似文献   

20.
Diagnosis of cardiovascular disease is currently limited by the testing modality. Serum tests for biomarkers can provide quantification of severity but lack the ability to localize the source of the cardiovascular disease, while imaging technology such as angiography and ultrasound can only determine areas of reduced flow but not the severity of tissue ischemia. Targeted imaging with ultrasound contrast agents offers the ability to locally image as well as determine the degree of ischemia by utilizing agents that will cause the contrast agent to home to the affected tissue. Ultrasound molecular imaging via targeted microbubbles (MB) is currently limited by its sensitivity to molecular markers of disease relative to other techniques (e.g., radiolabeling). We hypothesize that computational modeling may provide a useful first approach to maximize microbubble binding by defining key parameters governing adhesion. Adhesive dynamics (AD) was used to simulate the fluid dynamic and stochastic molecular binding of microbubbles to inflamed endothelial cells. Sialyl LewisX (sLex), P‐selectin aptamer (PSA), and ICAM‐1 antibody (abICAM) were modeled as the targeting receptors on the microbubble surface in both single‐ and dual‐targeted arrangements. Microbubble properties (radius [Rc], kinetics [kf, kr], and densities of targeting receptors) and the physical environment (shear rate and target ligand densities) were modeled. The kinetics for sLex and PSA were measured with surface plasmon resonance. Rc, shear rate, and densities of sLex, PSA, or abICAM were varied independently to assess model sensitivity. Firm adhesion was defined as MB velocity <2% of the free stream velocity. AD simulations revealed an optimal microbubble radius of 1–2 µm and thresholds for (>102 s?1) and (<10?3 s?1) for firm adhesion in a multi‐targeted system. State diagrams for multi‐targeted microbubbles suggest sLex and abICAM microbubbles may require 10‐fold more ligand to achieve firm adhesion at higher shear rates than sLex and PSA microbubbles. The AD model gives useful insight into the key parameters for stable microbubble binding, and may allow flexible, prospective design, and optimization of microbubbles to enhance clinical translation of ultrasound molecular imaging. Biotechnol. Bioeng. 2010;107: 854–864. © 2010 Wiley Periodicals, Inc.  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