Home>>Lab Reagents>> Detergents>>MEGA-10

MEGA-10

(Synonyms: N-decanoyl-N-Methylglucamine) 目录号 : GC44151

A detergent used to solubilize membrane proteins

MEGA-10 Chemical Structure

Cas No.:85261-20-7

规格 价格 库存 购买数量
500mg
¥619.00
现货

电话:400-920-5774 Email: sales@glpbio.cn

Customer Reviews

Based on customer reviews.

Sample solution is provided at 25 µL, 10mM.

产品文档

Quality Control & SDS

View current batch:

产品描述

MEGA-10 is a nonionic detergent that can be used to solubilize membrane proteins. [1] It has a critical micelle concentration (CMC) of 4.88 mM under no salt conditions, which decreases in the presence of sodium chloride. [2] MEGA-10 has been used to reconstitute amino acid transporters from rat liver plasma membrane vesicles into artificial phospholipid membranes.[3] It has also been used to solubilize the melibiose transport carrier from E. coli membranes and reconstitute it into liposomes.[4]

Reference:
[1]. Hildreth, J.E.K. N-D-Gluco-N-methylalkanamide compounds, a new class of non-ionic detergents for membrane biochemistry. Biochem. J. 207(2), 363-366 (1982).
[2]. Molina-Bolívar, J.A., Hierrezuelo, J.M., and Carnero Ruiz, C. Self-assembly, hydration, and structures in N-decanoyl-N-methylglucamide aqueous solutions: Effect of salt addition and temperature. J. Colloid. Interface Sci. 313(2), 656-664 (2007).
[3]. Quesada, A.R., and McGivan, J.D. A rapid method for the functional reconstitution of amino acid transport systems from rat liver plasma membranes. Partial purification of System A. Biochem J. 255(3), 963-969 (1988).
[4]. Hanatani, M., Nishifuji, K., Futai, M., et al. Solubilization and reconstitution of membrane proteins of Escherichia coli using alkanoyl-N-methylglucamides. J. Biochem. 95(5), 1349-1353 (1984).

Chemical Properties

Cas No. 85261-20-7 SDF
别名 N-decanoyl-N-Methylglucamine
化学名 1-deoxy-1-[methyl(1-oxodecyl)amino]-D-glucitol
Canonical SMILES OC[C@@H](O)[C@@H](O)[C@H](O)[C@@H](O)CN(C)C(CCCCCCCCC)=O
分子式 C17H35NO6 分子量 349.5
溶解度 1mg/mL in ethanol, 30mg/mL in DMSO or in DMF 储存条件 Store at -20°C
General tips 请根据产品在不同溶剂中的溶解度选择合适的溶剂配制储备液;一旦配成溶液,请分装保存,避免反复冻融造成的产品失效。
储备液的保存方式和期限:-80°C 储存时,请在 6 个月内使用,-20°C 储存时,请在 1 个月内使用。
为了提高溶解度,请将管子加热至37℃,然后在超声波浴中震荡一段时间。
Shipping Condition 评估样品解决方案:配备蓝冰进行发货。所有其他可用尺寸:配备RT,或根据请求配备蓝冰。

溶解性数据

制备储备液
1 mg 5 mg 10 mg
1 mM 2.8612 mL 14.3062 mL 28.6123 mL
5 mM 0.5722 mL 2.8612 mL 5.7225 mL
10 mM 0.2861 mL 1.4306 mL 2.8612 mL
  • 摩尔浓度计算器

  • 稀释计算器

  • 分子量计算器

质量
=
浓度
x
体积
x
分子量
 
 
 
*在配置溶液时,请务必参考产品标签上、MSDS / COA(可在Glpbio的产品页面获得)批次特异的分子量使用本工具。

计算

动物体内配方计算器 (澄清溶液)

第一步:请输入基本实验信息(考虑到实验过程中的损耗,建议多配一只动物的药量)
给药剂量 mg/kg 动物平均体重 g 每只动物给药体积 ul 动物数量
第二步:请输入动物体内配方组成(配方适用于不溶于水的药物;不同批次药物配方比例不同,请联系GLPBIO为您提供正确的澄清溶液配方)
% DMSO % % Tween 80 % saline
计算重置

Research Update

Solubilization of n-alkylbenzenes into decanoyl-N-methylglucamide (MEGA-10) solution; temperature dependence

Colloids Surf B Biointerfaces 2009 Feb 15;69(1):135-40.PMID:19150232DOI:10.1016/j.colsurfb.2008.12.011.

Solubilization of benzene, toluene, ethylbenzene, n-propylbenzene, n-butylbenzene, n-pentylbenzene, and n-hexylbenzene into micelles of decanoyl-N-methylglucamide (MEGA-10) was studied at 303.2, 308.2, 313.2, and 318.2K, where equilibrium concentrations of the above solubilizates were determined spectrophotometrically. The concentration of the above solubilizates remained constant below the critical micelle concentration (cmc) and increased linearly with an increase in MEGA-10 concentration above the cmc at each temperature above. The Gibbs free energy change of the solubilizates from aqueous bulk to their liquid solubilizate phase was evaluated from dependence of their aqueous solubility on alkyl chain length of the solubilizates, which leads to the DeltaG(CH0)(2) values (-3.60 to -3.38 kJ mol(-1)), the energy change per CH2 group of the alkyl chain with no strong temperature dependence. The first stepwise solubilization constant (K1) was evaluated from the slope for the change of solubilizate concentration vs. MEGA-10 concentration. The Gibbs free energy change (DeltaG(0,s)) for the solubilization decreased linearly with the carbon number of alkyl chain of the solubilizates, and the DeltaG(CH0)(2)(s) values (-2.71 to -2.54 kJ mol(-1)) obtained from the linearity showed a slight increase with temperature. The DeltaG(CH0)(2) values are less than the DeltaG(CH0)(2)(s) values, where the latter values clearly indicate that the location of alkyl chain is a hydrophobic micellar core. The fact is also supported by the absorption spectrum of the solubilized molecules. Temperature dependence of DeltaG(0,s) indicated that the solubilization is entropy-driven for the solubilizates with shorter alkyl chains, while it becomes enthalpy-driven for those with longer alkyl chains.

Solubilization of n-alkylbenzenes into decanoyl-N-methylglucamide (MEGA-10) solution

Langmuir 2008 Jan 1;24(1):15-8.PMID:18052401DOI:10.1021/la702820h.

Solubilization of benzene, toluene, ethylbenzene, n-propylbenzene, n-butylbenzene, n-pentylbenzene, and n-hexylbenzene into micelles of decanoyl-N-methylglucamide (MEGA-10) was studied, where equilibrium concentrations of the above solubilizates were determined spectrophotometrically at 303.2 K. The concentration of the above solubilizates remained constant below the critical micelle concentration (cmc) and increased linearly with an increase in MEGA-10 concentration above the cmc. The Gibbs free energy change of the solubilizates from the aqueous bulk to the liquid solubilizate phase was evaluated from the dependence of their aqueous solubility on the alkyl chain length of the solubilizates, which leads to -3.46 kJ mol-1 for DeltaG(0)(CH), the energy change per CH2 group of the alkyl chain. The first stepwise solubilization constant (K(overline)1 ) was evaluated from the slope of the change of solubilizate concentration versus MEGA-10 concentration. The Gibbs free energy change (DeltaG(0,s)) for the solubilization decreased linearly with the carbon number of the alkyl chain of the solubilizates, from which DeltaG(0,s)(CH2) as evaluated to be -2.71 kJ mol-1. The similar values above clearly indicate that the location of the alkyl chain is a hydrophobic micellar core, which is also supported by the absorption spectrum of the solubilized molecules.

Solubility properties of the alkylmethylglucamide surfactants

Biochim Biophys Acta 1990 Nov 2;1029(1):67-74.PMID:2223813DOI:10.1016/0005-2736(90)90437-s.

The critical micelle concentration (CMC) and the ability to solubilize and form vesicles from phospholipids are important criteria for the selection of a surfactant for reconstitution protocols. The CMC and its temperature dependence were determined for an homologous series of alkylmethylglucamides (MEGA-8, MEGA-9, MEGA-10). Each detergent was added continuously from a concentrated solution to a saline buffer with the environment-sensitive fluorescent probe ANS, held in a thermojacketed cuvette; ANS fluorescence increases at the CMC. The CMCs at 25 degrees C were 51.3, 16.0 and 4.8 mM for MEGA-8, MEGA-9 and MEGA-10. The free energy change for transfer to a micellar environment per -CH2- was -740 cal/mol, similar to other alkyl series. The CMCs decreased slightly with increasing temperature (T = 5-40 degrees C) for MEGA-9 and MEGA-10 while that of MEGA-8 was virtually insensitive to temperature in this range. MEGA-9 solubilization of egg PC in aqueous solutions was determined as a function of [PC] and temperature. The lamellar-micellar phase boundaries were determined by simultaneous 90 degrees light scattering and the resonance energy transfer using the headgroup labeled lipid probes NBD-PE and Rho-PE. The [MEGA-9] at solubilization was linear with [PC]; the MEGA-9 to egg PC ratio in the structures at optical clarity was 2.3 while the monomeric [MEGA-9] was 14.3 mM or slightly lower than the CMC at 25 degrees C. Solubilization of egg PC by MEGA-9 was somewhat more temperature-dependent than the CMC of this detergent. Vesicles formed from MEGA-9 tended to be multilamellar. MEGA-9 is clearly different from octyl glucoside, despite its chemical similarity, in terms of its temperature sensitivity and vesicle forming characteristics.

Complementary use of simulations and molecular-thermodynamic theory to model micellization

Langmuir 2006 Feb 14;22(4):1500-13.PMID:16460068DOI:10.1021/la052042c.

Molecular-thermodynamic descriptions of micellization in aqueous media can be utilized to model the self-assembly of surfactants possessing relatively simple chemical structures, where it is possible to identify a priori what equilibrium position they will adopt in the resulting micellar aggregate. For such chemical structures, the portion of the surfactant molecule that is expected to be exposed to water upon aggregate self-assembly can be identified and used as an input to the molecular-thermodynamic description. Unfortunately, for many surfactants possessing more complex chemical structures, it is not clear a priori how they will orient themselves within a micellar aggregate. In this paper, we present a computational approach to identify what portions of a surfactant molecule are hydrated in a micellar environment through the use of molecular dynamics simulations of such molecules at an oil/water interface (modeling the micelle core/water interface). The local environment of each surfactant segment is determined by counting the number of contacts of each segment with the water and oil molecules. After identifying the hydrated and the unhydrated segments of the surfactant molecule, molecular-thermodynamic modeling can be performed to predict: (i) the free-energy change associated with forming a micellar aggregate, (ii) the critical micelle concentration (CMC), and (iii) the optimal shape and size of the micellar aggregate. The computer simulation results were found to be sensitive to the atomic charge parameters utilized during the simulation runs. Two different methods of assigning atomic charges were tested, and the computer simulation and molecular-thermodynamic modeling results obtained using both sets of atomic charges are presented and compared. The combined computer simulation/molecular-thermodynamic modeling approach presented here is validated first by implementing it in the case of anionic (sodium dodecyl sulfate, SDS), cationic (cetyltrimethylammonium bromide, CTAB), zwitterionic (dodecylphosphocholine, DPC), and nonionic (dodecyl poly(ethylene oxide), C12E8) surfactants possessing relatively simple chemical structures and verifying that good predictions of CMCs and micelle aggregation numbers are obtained. In the case of C12E8, the challenges and limitations associated with simulating a single, polymeric E8 moiety at the oil/water interface to model its behavior at the micelle/water interface are discussed. Subsequently, the combined modeling approach is implemented in the case of the anionic surfactant 3-hydroxy sulfonate (AOS) and of the nonionic surfactant decanoyl-n-methylglucamide (MEGA-10), which possess significantly more complex chemical structures. The good predictions obtained for these two surfactants indicate that the combined computer simulation/molecular-thermodynamic modeling approach presented here extends the range of applicability of molecular-thermodynamic theory to allow modeling of the micellization behavior of surfactants possessing more complex chemical structures.

Physicochemical Studies on the Interaction between N-Decanoyl-N-methylglucamide and Bovine Serum Albumin

Biomacromolecules 2007 Aug;8(8):2497-503.PMID:17630693DOI:10.1021/bm0704121.

The protein-surfactant system constituted by bovine serum albumin (BSA) and N-decanoyl-N-methylglucamide (MEGA-10) has been studied by using surface tension, steady-state fluorescence, and dynamic light scattering measurements. It was found that the presence of protein delays the surfactant aggregation, which was interpreted as a sign of binding between surfactant and protein. Binding studies were carried out by two different methods. First, a treatment based on surface tension measurements was used to obtain information on the number of surfactant molecules bound per protein molecule under saturation conditions. Second, the binding curve for the BSA/MEGA-10 system was determined by examining the behavior of the intrinsic BSA fluorescence upon the surfactant addition. Both approaches indicate that the binding process is essentially cooperative in nature. The results of the aggregation numbers of MEGA-10 micelles, as well as those of resonance energy transfer from tryptophan residues to 8-anilinonaphthalene-1-sulfonate, corroborate the formation of micelle-like aggregates of surfactants, smaller than the free micelles, adsorbed on the protein surface. The dynamic light scattering results were not conclusive, in the sense that it was not possible to discriminate between protein-surfactant complexes and free micelles. However, the overall results suggest the formation of "pearl necklace" complexes in equilibrium with the free micelles of the surfactant.