The largest database of trusted experimental protocols

Methylamine

Methylamine is a simple organic compound with the chemical formula CH3NH2.
It is a colorless, flammable gas with a fishy odor, commonly used in the synthesis of various pharmaceuticals, agrochemicals, and other industrial chemicals.
Methylamine plays a critical role in the production of important substances like dimethylamine, trimethylamine, and various methylated derivatives.
Researchers studying Methylamine can utilize PubCompare.ai's AI-driven platform to easily locate and compare experimental protocols from literature, preprints, and patents, enhancing the reproducibility and accuracy of their work.
This powerful comparison tool can help uncover the best methods and products for Methylamine-related projects, optimiziing the research process and driving innovation in this field.

Most cited protocols related to «Methylamine»

Consider a macromolecular system of n (nonhydrogen) atoms with Cartesian coordinates. To rapidly evaluate the energy of a particular configuration of the system (including hydrogens), we will decompose the system into a collection of distinct chemical groups, {Ai}, consisting of atoms for which the protonation state is unknown and a set P, the part of the system for which there is assumed to be no uncertainty regarding its protonation state.
The decomposition proceeds as follows: implicitly break all bonds between 4-coordinated alkane sp3 carbon atoms and collect the resulting connected (bonded) groups of atoms. For proteins, this will leave the backbone intact, isolate the alkane carbons, and produce a collection of m-methylamide (Asn, Gln), thiomethanol (Cys), methylimidazoles (His), methylguanidinium (Arg), methyl carboxylic acids (Asp, Glu), methanol (Ser, Thr), indole (Trp), methylphenol (Tyr) and methylbenzene (Phe), methylamine (Lys), and thioether (Met) groups. A special case disconnection of the standard termini will produce a methyl amine (N terminus) and a methyl carboxylic acid (C terminus). Solvent and disconnected ions are considered to be separate groups. Collect the backbone and isolated alkane atoms into a set, P, the “known” portion of the system. The remaining atoms in the chemical groups are collected (by connectivity) into m sets, {Ai}, the sets of the atoms for which there is uncertainty with respect to their protonation geometry, tautomer, or ionization state. This decomposition procedure assumes that alkane carbons and the protein peptide backbone have a known protonation state. In principle, any partitioning method can be used by Protonate3D provided that (relatively) apolar bonds are used to divide the system. The reason for this has to do with the thermodynamic approximations and the calculation of partial charges (which will be described later).
The hydrogen atoms of the heavy atoms of P (the “known” atoms) are added at standard bond lengths and angles according to the hybridization state of the atoms; for example, the backbone nitrogen in nonproline peptide bonds is given one hydrogen in the peptide plane; the Cα of nonglycine residues is given one hydrogen placed in an ideal tetrahedral geometry; sp3 carbons with two heavy neighbors (e.g., Cβ of Glu) are given two hydrogens placed at ideal tetrahedral geometry; terminal methyls are given three hydrogens in tetrahedral geometry in staggered conformation with respect to their (necessarily) alkane carbon neighbors. Henceforth, P will denote the hydrogen augmented set of atoms in the “known” part of the macromolecule.
For each chemical group Ai, we generate a finite collection Si = {Ai1,Ai2,…} of states consisting of the heavy atoms, flipped states, and all rotamer, tautomer, and ionization/protonation combinations of hydrogen atoms (see Fig. 1). In general, the states of chemical groups are generated according to a parameter file containing definitions of each chemical group and all of their topological tautomer and ionization states. The parameter file also contains, for each state, a tautomer strain energy (to provide for tautomer preferences). Rotamer (conformational) strain energy of each state is also considered and generated from force field parameter files such as OPLS-AA18 by applying the dihedral energy terms to the fragment geometry (as though still connected to P) and the intrafragment van der Waals energy terms (interfragment energies are handled by the matrix formulation of Eq. (1), later).
For proteins, the sp3 carbon atoms with two heavy neighbors are given hydrogens in a similar manner to the carbons of P; sp2 carbon atoms with one heavy neighbor (e.g., aromatic carbons) are given one hydrogen at standard bond lengths and angles in the π system plane. Primary amides are given two hydrogens at standard planar geometry; planar nitrogen atoms with two heavy neighbors and one hydrogen has that hydrogen placed in-plane at standard bond lengths and angles. The polar hydrogens and terminal methyls are given hydrogens appropriate to their ionization state and hybridization at standard bond lengths and angles. The dihedral combinations are determined according to the chemical type of the heavy atom: hydrogens in hydroxyls and thiols are sampled at 60° dihedral increments starting at a staggered rotamer; phenol hydrogens and other conjugated hydroxyls are sampled at 30° dihedral increments starting at an in-plane rotamer; methyls and primary amines are sampled at 60° dihedral increments starting at an extended conformation; hydrogens on other terminal atoms are given similar geometries. The anionic state of phenols, alcohols, thiols, and indoles are generated in addition to the neutral forms. The flip states of terminal amides, sulfonamides, and phosphonamides are generated. The anionic state and both neutral tautomers of carboxylic acids are generated (with the hydrogen cis to the carbonyl oxygen). Primary amines are generated in neutral and cationic forms and dihedral angles sampled at 60° increments starting at a staggered rotamer. Imidazoles are generated in anionic, cationic and two neutral tautomers (HID and HIE) as well as in flipped states (for a total of eight states). The states neutral of guanidines consist of all planar tautomers and rotamers. Water states consist of ∼500 rigid body orientations and isolated metals are given appropriate ionization states for groups I and II and a collection of ionization states from {+1,+2,+3} for transition metals under the assumption of zero ionization potential.
Thus, each Aij consists of an all-atom chemical group with an appropriate ionization state, the heavy atoms, all of its hydrogen atoms in reasonable geometry and has an associated internal energy, sij, consisting of the sum of its conformational and tautomeric energy. Figure 1 depicts a hypothetical fixed part P (with known protonation state and geometry) of a macromolecular system and three chemical groups each with a collection Si of alternative protonation states; A1 has four alternative states, A2 has two states, and A3 has three states.
To represent the state ensemble of the system, arrange all of the individual chemical group states in all of the {Si} into single state list, S, divided into contiguous blocks corresponding to the {Si}, each of length mi = |Si|.
The first block of m1 elements in the list are the states of chemical group 1, the next block of m2 elements in the list are the states of group 2, and so on. (The reason for this arrangement will become clear shortly.) A configuration of the entire system consists of a selection of exactly one particular state from each block associated with a chemical group. Thus, there are a total of m1 × m2 × m3 × … configurations of the system. In typical proteins, the number of configurations exceeds 10100. A binary vector x of length equal to the length of the list S conveniently encodes a configuration, with a value 1 denoting the selection of an individual state. For example, in Figure 1, the vector x = (0,1,0,0,1,0,0,0,1) denotes the configuration state 2 from group 1, state 1 from group 2, and state 3 from group 3; to see this, introduce dividers into x corresponding to the blocks: x = (0,1,0,0 | 1,0 | 0,0,1), so that the position of the 1 value within each block (counting from the left) indicates the number of the state within the group. Admissible, or permitted, configuration vectors, x, have the property that there is exactly one 1 value in each block corresponding to a chemical group; this means that an admissible configuration vector encodes a definite single state for each chemical group. This constraint giving rise to the admissible configuration vectors is called the unary constraint, inspired by unary (base 1) notation of numbers in which “1” = 1, “10” = 2, “100” = 3, “1000” = 4, “10,000” = 5, and so on.
Suppose that we are given a pairwise interaction energy function f(i,j), for atoms i and j (e.g., Coulomb's law or a Lennard-Jones van der Waals potential), without loss of generality, we will assume that f(i,i) is well defined (e.g., for Coulomb's law, f(i,i) = 0). If X and Y are two disjoint sets of atoms (e.g., two chemical states), then the interaction energy between X and Y is
Form a matrix U with entries equal to the interaction energy of the various chemical group states in the list S. We will take the interaction energy between two states of the same chemical group to be zero. For notational convenience, let I(k) denote the chemical group to which state k belongs. Thus, the matrix U will have Uij = f(Ai,Bj) if I(i) ≠ I(j) and 0 otherwise. Form a vector u with entries ui = f(P,Ai) + si, the interaction energy between a chemical group state and the known part of the protein, P, and the internal energy of the state, si (to be described later). Let u0 = f(P,P)/2, the (constant) internal interaction energy of the known part of the protein P. With this matrix notation, we can write the total energy of a particular configuration encoded by admissible binary vector, x, compactly (and efficiently) with
Thus, the total energy of a configuration of the system specified by x can be evaluated by a multidimensional quadratic form. If all of the values of u and U are calculated in advance, then a matrix–vector multiplication and two inner products are all that is required to evaluate the total energy for any arbitrary configuration of the system. Finding the optimal configuration of the system now is a matter of finding the smallest value of the quadratic form E over all binary vectors x satisfying the unary constraint; this optimization problem is called the “Unary Quadratic Optimization” problem.
Postponing the details of the energy model, the algorithmic structure of Protonate3D is (a more detailed set of steps is given at the end of this section):
The addition of many (more than 20) water molecules (each with ∼500 orientations) becomes impractical. As a result, most of the water molecules are typically left out of the preceding steps and oriented afterward. This is done by orienting the waters one by one proceeding from the water in the strongest electrostatic field (of the protein and previously oriented waters) to the weakest. The selection of water molecules to include in the main calculation is left to the user—typically, water molecules near the sites of interest are treated in the main calculation.
The Unary quadratic optimization algorithm used by Protonate3D proceeds as follows. First, a dead-end elimination14 (link) procedure is applied to eliminate states that cannot possibly be part of the optimal solution. This has the effect of reducing the dimensions of the U matrix and u vector of the quadratic energy function in a provably correct way. Suppose, elements r and s of the list S belong to the same chemical group X; if (where the sum extends over all chemical groups Y different from X) we can eliminate state r. The dead-end elimination criterion, when satisfied, eliminates r because no matter what state assignment is made, some state X, different from r, will result in a lower energy. This criterion is applied repeatedly until no more elimination is possible. Typically, the majority of the configurations are eliminated a priori, but it is still not practical to conduct a brute force search over the remaining configurations.
In an effort to speed up the state space search to follow, a “Mean Field Theory” calculation is performed to produce a Boltzmann distribution over all of the remaining individual chemical group states. This results in an estimate of the probability of each state in the Boltzmann-weighted ensemble of configurations. Briefly, the state probabilities pk are determined by solving the nonlinear equation. where p is the probability vector; U and u are as in Eq. (1); ek is a vector of all zeros and a single 1 at position k; and β = 1/kT. The nonlinear equation can be solved efficiently by successive feedback iteration. These probabilities, p, are the population probabilities of the individual states under the assumption that each state feels the Boltzmann weighted average interactions of the other states. The vector p is used as a heuristic state priority in the subsequent search over states; the idea is to investigate high mean field probability states first under the assumption that they will lead to low energy configurations of the entire system (an approximate best-first search). The mean field probabilities, p, only affect the run-time of the state-space search and not its correctness; moreover, the energy of a system is evaluated using Eq. (1), which does not depend on p. The value of β must be chosen carefully to guarantee the uniqueness of p; in general, the solutions to Eq. (3) depend on the starting p vector. However, for certain values of β, the solution will be independent of the starting point (see the Appendix) and consequently p can be initialized with a uniform distribution on the states of each chemical group.
Finally, a recursive tree search is conducted over all admissible binary vectors, x, to locate the lowest energy state as calculated by Eq. (1) (which provides for rapid evaluation of energies). The performance of the search depends critically on the ability to prune the search space without loss of correctness. At any given point in the search, some of the elements of x, corresponding to some set of groups, G, will be assigned and others are yet to be assigned (with zero values). A lower bound, L(x), on the minimum energy of the system assuming the assigned part of x is
If this lower bound value exceeds the energy of the best energy determined thus far, then no further search of configurations containing the assigned part of x is required, thereby pruning the search tree and bypassing the examination of descendant configurations. During the recursive search, trial elements of the unassigned portion of x are made in decreasing order of the mean field probability. This greatly improves the pruning performance of the lower bound because the likelihood of visiting the best configurations first is increased. Moreover, premature termination of the search will produce the best solution with high probability.
The pseudocode for the recursive tree search procedure is given in Figure 2.
We now turn to the energy model for the macromolecular system. We will use an energy model that contains van der Waals repulsion, Coulomb electrostatic, and Generalized Born implicit solvation energies. Use of the Poisson-Boltzmann Equation (PBE) was not attempted because it was felt that the run-time would be prohibitively long for large systems, requiring at least one PBE solution per state. The van der Waals and Coulomb functional forms terms are pairwise and fit neatly into the quadratic form of Eq. (1); however, the Generalized Born model is not a two-body potential and certain approximations will be used to reformulate it into an effective two-body potential. In addition, because the number of particles may change upon ionizing a chemical group, we must introduce free energy terms related to group titration (because potential energies cannot be compared for systems with different numbers of particles).
Each atom of the system, whether in the known part, P, or in one of the group states {Aij} has associated van der Waals radius, van der Waals well depth parameters, as well as a partial charge. The van der Waals parameters and partial charges are permitted to depend on the particular tautomer, rotamer, or ionization state of each chemical group. In the interests of efficiency, we impose the requirement that the van der Waals parameters and partial charge assignments of one chemical group do not depend on the particular state selection of another chemical group. In particular, we require that the partial charge model be a nonpolarizable charge model (see the titration theory, later). The decomposition of the system along apolar bonds is done to reduce the potential adverse impact of these independence requirements.
Protonate3D uses a slightly modified version of MMFF9419 partial charges because (a) the MMFF94 charge model is based on fixed (topological) bond charge increments; (b) the chemical contexts for atom types in MMFF94 do not cross sp3 carbon atoms; (c) the bond charge increment between sp3 carbon bonds is zero (a purely apolar bond); and (d) MMFF94 supports general organic compounds. The slight modification to the MMFF94 charge model is that the normal zero bond charge increment between alkane hydrogens and carbons was replaced with a bond charge increment of 0.08 electrons, in better agreement with protein force field partial charges such as AMBER.20 Protonate3D uses Engh–Huber21 van der Waals parameters; however, hydrogens on oxygen and nitrogen are taken to have zero van der Waals radius, consistent with OPLS-AA. Coulomb's law is used for electrostatic interactions and special form of van der Waals interaction is used: only the repulsive part of the van der Waals interaction energy is modeled (although, the standard Lennard-Jones functions with the attractive term are not precluded). The special functional form is 800εij (1 − r/Rij),3 (link) where r < Rij is the interatomic separation, Rij is the sum of the van der Waals radii, and εij is the geometric mean of the van der Waals well depth parameters for the two interacting atoms. Because of the 800 factor derived from a series expansion, this functional form lies in between the 12-6 and 9-6 Lennard-Jones functions at distances below the optimal interaction distance and approximates the 12-6 form well near the energy minimum. Because the OPLS-AA van der Waals parameters for polar hydrogen atoms are zero, the van der Waals terms are used by Protonate3D to handle side-chain “flip” states; the special form was used largely to mimic the sphere overlap test of Reduce.7 (link) The elements of U matrix and u vector are populated by a straightforward application of the pairwise formulae given previously.
Protonate3D uses a modified version of the Generalized Born/Volume Integral (GB/VI) formalism22 (link) for implicit solvent electrostatics (although other Generalized Born models are not precluded):
In this equation, ε is the dielectric constant of the interior of a solute, εsol is the dielectric constant of the solvent, {γi} are (topological) atom-type-dependent constants that account for nonpolar energies including cavitation and dispersion using an inverse sixth-power integral instead of surface area, {Ri} are (topological) atom-type-dependent solvation radii, κ is the Debye ionic screening parameter that depends on salt concentration, {qi} are the atomic partial charges, {Bi} are the Born self-energies (inversely proportional to the Born radii), which are estimated with a pairwise sphere approximation23 to the solute cavity, and rij denotes the distance between atoms i and j. Were it not for the {Bi}, the GB/VI equations would be a pairwise potential; however, because the Bi of a particular atom i depends on the state assignment of atoms in other chemical groups with possibly unknown state, we must calculate a set of {Bi} that (a) remain fixed despite the protonation state of other groups and (b) reasonably preserve the GB/VI energy values.
Consider an atom k in the system (whether in P or in some state Aij). The contribution to Bk from all of the other atoms in the system will fall as the sixth power in the integrand of Eq. (7). Thus, atoms far away from k will contribute little, no matter if they are in some other group with unknown state. The various states in the system differ only in the position or absence of hydrogen atoms, which contribute relatively little to the volume integral (because of their small solvation radius); thus, the bulk of the states' contribution (from the heavy atoms) will be accurate no matter which state is selected. In any event, the approximation to the volume integral in the GB/VI is a pairwise summation of the form for a specific function22 (link) V. To minimize the impact of the hydrogen positions of the unknown states, Protonate3D uses a separate mean field approximation to the volume integrals. A separate U matrix and u vector is created containing only the van der Waals repulsion terms, the states' internal strain energies, and the pH-dependent isolated group titration energies (see later). For each separate group state, the mean field equation of Eq. (3) is then solved to produce a set of state probabilities p. Each atom in each group state as well as the known part P is given the probability of its chemical group state, or 1 if the atom is in P. The Born factors are then calculated with resulting in a mean field approximation to the Born factors that takes steric, rotamer/tautomer, and isolated group pKa free energies into account. This approximation works very well in practice; indeed, one can argue it is in some sense superior to the original in that it takes some protonation state flexibility into account. It should be noted that some GB implicit solvent models do not include hydrogens in the volume integration24 (link); consequently, we believe that our calculation of the mean field Born factors is eminently reasonable. In this way, we approximate the three-body GB/VI model with a close pairwise model more suited for the quadratic form of Eq. (1).
It remains to deal with the pH-dependent free energy of ionization of the chemical groups that must be included in the calculation. Consider the free energy, a, of the reaction PAH → PA + H+, where AH is an acidic group bound (possibly covalently) to a macromolecule P. Our approach is to introduce a thermodynamic cycle linking the reaction to the isolated group reaction AH → A + H+, whose free energy will be assumed to be known. In the covalent case, we consider the thermodynamic cycle in which a = b + c + d. If the pKa of the reaction HAH → HA + H+ is known (say from experiment), then for a given pH, we have that c = −kT (log 10) (pH−pKa), where k is Boltzmann's constant and T is the temperature of the system. Because the (vertical) reaction equation H2 + PAH → PH + HAH is balanced and, by construction, E = ECOUL + ESOL is the free energy of charging and solvating the system, we may simply write
The case of a noncovalently bound group AH near a macromolecule P is simpler in that the H2 molecule is not required to balance the equation and, in this case,
We shall deal with the noncovalent case first, because it is simpler and provides insight into the covalent case. The noncovalent d is similar to b resulting in and using the fact that E(A + B) = E(A) + E(B) we have that
The superscript iso is used to signify that the E is calculated for the isolated AH and A systems (i.e., calculated with Born factors derived from the isolated system, ignoring P). These iso superscripted quantities involve only a small number of atoms—the atoms of AH and A—and direct evaluations of E are used to calculate the required energy. The iso superscripted quantities are included directly in the u vector of Eq. (1) for the corresponding group state so that b + d is simply a difference of configuration energies.
With a similar line of reasoning as in the noncovalent case, we find that as a result of cancellations of E(PH) and E(H2), for the covalent case and, as before, the iso superscripted quantities can be calculated directly (because few atoms are involved) and included in the u vector. In practice, the distinction between covalent and noncovalent groups makes only a small difference—on the order of 0.5 kcal/mol (∼2% error) for ionic species. A small correction to the experimental isolated pKa values for covalently bound species can account for most of this difference. In any event, the static nature of the entire calculation and the approximations inherent in a Generalized Born model will in all likelihood overshadow any lack of distinction between the cases.
The free energy c = −kT (log 10) (pH − pKa) remains to be included in Eq. (1). Consider a polyprotic species AHn with pKa values pKi, corresponding to AHi → AHi−1. The free energy of the reaction AHi → AHi−1 + H+ is then ΔGi = −kT log 10 (pH − pKi). If we assign we will have that ΔGi = GiGi−1; thus, we can incorporate the Gi values into the relevant u vector entries for each acidic chemical group state with i titratable protons. The reasoning for the b and d quantities generalizes to polyprotic species and multiple-site titration straightforwardly, because of the overall pairwise nature of the energy terms that make up the effective configuration energy.
We now summarize the Protonate3D procedure:
This brings to a close the exposition of the Protonate3D methodology. Protonate3D was implemented in the Scientific Vector Language of the Molecular Operating Environment25 version 2006.08. Computational experiments were conducted on a 2 GHz Pentium IV processor running Microsoft Windows.
Publication 2008
To validate that our medium, henceforth “MP”, compared well to other formulations currently used to grow M. extorquens, growth rate was compared for AM1Δcel growing on methylamine and methanol, as well as AM1Δcel and PA1Δcel on succinate, in MP and four other media. The first medium we tested was our historically used variant-Hypho (aged for over four weeks). The second and third media tested were phosphate-buffered media that differed in initial pH (second media: initial pH = 6.7 for growth on multi-carbon compounds, henceforth “Phosphate-multi-C”; third media: initial pH = 7.1 for growth on C1 compounds, here “Phosphate-C1” [39] ). The rationale behind testing different pH levels was to partially counter the tendency that growth on multi-C substances increases pH, whereas the opposite is seen for C1 compounds. The final medium we compared, Choi medium [40] , is a Methylobacterium medium developed to aid poly-β-hyroxybutyric acid (PHB) production and has an exceptionally metal-rich formulation; total trace metals are in the mM range instead of the µM range. A comparison of the concentrations of main components of each of these media is given in Table 2.
On C1 compounds, strains grown in MP medium grew faster than in all other media (Fig. 3). With methylamine as the substrate, the growth rate on MP was estimated to be 11% faster than on our older variant-Hypho, and 15% faster than on Phosphate-C1 medium (all p-values <1×10−6). With methanol as the substrate, due to evaporation, the cultures did not achieve an OD over 0.1 and could not be fit over the same range of OD values; however, when fit over an OD range of 0.01–0.07, the MP medium was estimated to be 7% and 17% faster than on variant-Hypho and Phosphate-C1 (p-values <1×10−6), respectively. We did not make comparisons to the Choi media as it produced data that was too noisy for meaningful analysis (Fig. S3). Although the Choi medium did appear to have growth rates similar to the other media tested, the large concentration of unchelated metals in Choi medium formed dense precipitates on the bottom of the wells, making it difficult to set a well’s blank value and causing erratic OD measurements throughout the growth period (Fig. S3). For this reason, meaningful quantitative comparisons could not be made and we concluded that Choi medium could not be used for growth rate measurements in microtiter plates (Fig. S3).
On succinate, strains grown in MP medium performed as well as or better than the other media we tested (Fig. 3, Fig. S4). For AM1Δcel growing on succinate, the mean growth rate was estimated to be 6% faster with MP as compared to Hypho medium (p = 1.74 ×10−8), while for PA1Δcel there was a smaller but still significant improvement of 1.7% compared to Hypho medium (p = 0.01). Although the Phosphate-multi-C medium initially appeared to grow similarly to variant-Hypho or MP, its growth rate noticeably slowed as OD increased in a clear violation of the exponential growth model (Fig. S4) and so it was not quantitatively compared to the other media.
Publication 2013
Acids Carbon Growth Substances Metals Methanol methylamine Methylobacterium Phosphates Poly A Strains Succinate Vision
Lipids were extracted as previously described (Guan et al., 2010 (link)) with minor modifications. Briefly, cells (25 OD600 units) were resuspended in 1.5 ml of extraction solvent (ethanol, water, diethyl ether, pyridine, and 4.2 N ammonium hydroxide [15:15:5:1:0.018, vol/vol]). A mixture of internal standards (7.5 nmol of 17:0/14:1 PC, 7.5 nmol of 17:0/14:1 PE, 6.0 nmol of 17:0/14:1 PI, 4.0 nmol of 17:0/14:1 PS, 1.2 nmol of C17:0-ceramide, and 2.0 nmol of C8-glucosylceramide) and 250 μl of glass beads was added, the sample was vortexed vigorously (multitube vortexer; Labtek International, Christchurch, New Zealand) at maximum speed for 5 min and incubated at 60°C for 20 min. Cell debris was pelleted by centrifugation at 1800 × g for 5 min, and the supernatant was collected. The extraction was repeated once, and the supernatants were combined and dried under a stream of nitrogen or under vacuum in a Centrivap (Labconco, Kansas City, MO). The sample was divided into two equal aliquots. One half was used for ceramide and sphingolipid analysis, in which we performed an extra step to deacylate glycerophospholipids using monomethylamine reagent (methanol, water, n-butanol, methylamine solution [4:3:1:5, vol/vol]; Cheng et al., 2001 (link)) to reduce ion suppression due to glycerophospholipids in sphingolipid detection. For desalting, both lipid extracts were resuspended in 300 μl of water-saturated butanol and sonicated for 5 min. We added 150 μl of LC-MS–grade water, and samples were vortexed and centrifuged at 3200 × g for 10 min to induce phase separation. The upper phase was collected. Another 300 μl of water-saturated butanol was added to the lower phase, and the process was repeated twice. The combined upper phases were dried and kept at −80°C until analysis. For glycerophospholipid and sphingolipid analysis by electrospray ionization-MS/MS, lipid extracts were resuspended in 500 μl of chloroform:methanol (1:1, vol/vol) and diluted in chloroform:methanol:water (2:7:1, vol/vol/vol) and chloroform:methanol (1:2, vol/vol) containing 5 mM ammonium acetate for positive and negative mode, respectively. A Triversa Nanomate (Advion, Ithaca, NY) was used to infuse samples with a gas pressure of 30 psi and a spray voltage of 1.2 kV on a TSQ Vantage (ThermoFisher Scientific, Waltham, MA). The mass spectrometer was operated with a spray voltage of 3.5 kV in positive mode and 3 kV in negative mode. The capillary temperature was set to 190°C. MRM-MS was used to identify and quantify lipid species as previously described (Guan et al., 2010 (link)). Data were converted and quantified relative to standard curves of internal standards that had been spiked in before extraction. Two independent biological replicates were analyzed, each of which comprised up to six technical replicates.
For glycerophospholipid and sphingolipid analysis by high-resolution mass spectrometry, we used a Q Exactive Hybrid Quadrupole-Orbitrap Mass Spectrometer (Thermo Scientific). Lipid extracts were dissolved in 500 μl of chloroform:methanol (1:1, vol/vol) and diluted two times in 5 mM ammonium acetate in chloroform:methanol:water (2:7:1, vol/vol/vol) for both positive- and negative-ion-mode mass analysis. Samples were analyzed by direct infusion. Positive-ion-mode analysis was performed using scan range m/z = 650–800 (for monitoring PC and PE) and 540–750, using lock mass 588.4471 (for ceramides). Negative-ion-mode analysis was performed using scan range m/z = 700–850 (for monitoring PI and PS) and 550–1150 (for complex sphingolipids). The acquisition was set to 100 scans. Spectra were acquired using mass resolution of 280,000 and automatic gain control at 3e6. Lipid species were identified according to their m/z, and their abundance was calculated by their signal intensities relative to internal standards (17:0/14:1 PC for PC and PE, 17:0/14:1 PI for PI, 17:0/14:1 for PS, C8-glucosylceramide for complex sphingolipids, and C17 ceramide for ceramides).
Publication 2014
For the two pairs of antibodies where crystal structures were available, those were used as starting structures for metadynamics simulations. All structures were prepared in MOE (Molecular Operating Environment) (40 ) using the Protonate3D (41 (link)) tool. The C-termini of the antibodies were capped with N-Methylamine (NME). With the tleap tool of the AmberTools16 (38 ) package, the two systems were soaked into cubic water boxes of TIP3P water molecules (42 (link)) with a minimum wall distance to the protein of 10 Å. Parameters for all antibody simulations were derived from the AMBER force field 14SB (43 (link)). Each system was carefully equilibrated using a multistep equilibration protocol (44 (link)). To achieve an extensive but efficient exploration of the conformational space, well-tempered metadynamics simulations were performed using GROMACS (45 (link)), i.e., plumed 2 (46 (link)) software package. In metadynamics simulations a history-dependent bias potential is built based on Gaussian functions, which are deposited on the potential energy surface at already sampled conformations (47 (link)). This leads to an accelerated sampling allowing the system to escape deep energy minima. Well-tempered metadynamics (48 (link)) adapts the height of the Gaussian functions with simulation time. Various collective variables (CV) have been tested to achieve a better description of the conformational space. The most efficient CVs for our systems were found to be linear combinations of sine and cosine of the ψ torsion angles (49 (link)) of the CDR-H3 and CDR-L3 loops, which were calculated with functions MATHEVAL and COMBINE implemented in plumed 2 (46 (link)). As discussed previously, the ψ torsion angle captures conformational transitions comprehensively (50 (link)–52 (link)). The decision to include the CDR-L3 loop is based on previously observed structural correlation of the CDR-L3 and CDR-H3 loop (53 (link)). The height of the Gaussian was determined according to minimal distortion of the antibody systems, resulting in 10.0 kcal/mol for the antibodies with structural information and 2.0 kcal/mol for the Fab 246 and Fab 249. Gaussian deposition occurred every 1,000 steps and a biasfactor of 10 was used. For all 6 antibodies we collected for each starting structure 1 μs by metadynamics simulations. The resulting trajectories were aligned to the entire Fv and clustered in cpptraj (38 , 39 (link)) using the average-linkage hierarchical clustering algorithm. The Cα-RMSD of the CDR-H3 loop was used as distance metric and the same cutoff criterion was applied for each pair of antibodies. The choice of the distance cutoff is optimized to obtain a broad cluster distribution within Principle Component Analysis (PCA) space for each system. For the Fab 246 and Fab 249 antibody pair we chose 200 cluster representatives each to compensate for the uncertainty introduced by using modeled structures. The resulting cluster representatives for all systems were equilibrated and simulated for 100 ns using classic MD as implemented in the AMBER16 simulation package (38 ). Molecular dynamics simulations were performed in an NpT ensemble using pmemd.cuda (38 , 54 (link), 55 (link)). Bonds involving hydrogen atoms were restrained by applying the SHAKE (56 (link)) algorithm, allowing a time step of 2.0 fs. Atmospheric pressure of the system was preserved by weak coupling to an external bath using the Berendsen algorithm (57 (link)). The Langevin thermostat (58 (link)) was used to maintain the temperature at 300 K during simulations.
The obtained MD trajectories for each system were analyzed with PCA and the time-lagged independent component analysis (tICA) (59 (link)) of the Cα CDR-H3 loop atoms using the python library PyEMMA 2 (60 (link)) and employing a lag time of 5 ns. tICA can be used as a dimensionality reduction method and is a technique to find the slowest-relaxing degrees of freedom. Compared to PCA, which leads to high-variance linear combinations of the input degrees of freedom, tICA shows high-autocorrelation linear combinations of the input degrees of freedom (61 (link), 62 (link)). The tICA space was used for clustering to generate microstates that build the basis for a MSM. The aim of the Markov-state models is to define kinetically relevant states, to estimate the transition times between them and to quantify the probability of the states (63 (link)). Thus, kinetics were estimated by constructing a Markov-state model (63 (link)) employing a lag time of 5 ns using PyEMMA 2. We chose the lag time according to the implied timescale plot, which shows an approximately constant behavior of the estimated timescales at lag times over 5 ns (64 (link), 65 (link)). We used k-means clustering (60 (link)) to define 200 microstates and the PCCA+ algorithm (66 (link)) to calculate macrostates and estimate their according representative structures. PCCA+ is a spectral clustering method, which discretizes the sampled conformational space based on the eigenvectors of the transition matrix.
Publication 2018
The atomic polarizabilities are listed for each AMOEBA atom type in Table 1. The values are the same as those derived by Thole62 except for aromatic carbon and hydrogen atoms which have been systematically refined using a series of aromatic systems, including a small carbon nanotube (see Table 3). The molecular polarizabilities computed using the current model are compared to experimental values for selected compounds in Table 2. Reducing the damping factor from Thole’s original value of 0.567 to AMOEBA’s 0.39 is critical to correctly reproducing water cluster energetics.18 On the other hand, AMOEBA’s greater damping leads to a slight systematic underestimation of molecular polarizabilities. However, given the simplicity of the model, the agreement is generally satisfactory for both average polarizabilities and their anisotropies. As described above, polarization groups are defined for purposes of treating intramolecular polarization. Typically a functional group is treated as a single polarization group. For example, methylamine is a group by itself, while ethylamine has two groups: –CH2NH2 and CH3–. The groups are specified in AMOEBA parameter files in the following format: “polarize A α 0.390 B C”, where α is the polarizability for atom type A; 0.39 is the damping coefficient in Eq (2); B and C are possible bonded atom types that belong to the same polarization group as atom type A.
The permanent atomic multipoles were derived for each molecule from ab initio QM calculations. Ab initio geometry optimization and a subsequent single-point energy evaluation were performed at the MP2/6-311G(1d,1p) level using Gaussian 03.78 For small molecules with less than six heavy atoms, Distributed Multipole Analysis (DMA v1.279 ) was used to compute the atomic multipole moments in the global frame using the density matrix from the QM calculation. Next, the TINKER POLEDIT program rotates the atomic multipoles into a local frame and extracts Thole-based intramolecular polarization to produce permanent atomic multipole (PAM) parameters. Thus, when the AMOEBA polarization model is applied to the permanent atomic moments, the original ab initio-derived DMA is recovered. Finally, the POTENTIAL program from the TINKER package is used to optimize the permanent atomic multipole parameters by fitting to the electrostatic potential on a grid of points outside the vdW envelope of the molecule. The reference potential for the fitting step is typically derived from a single point calculation at the MP2/aug-cc-pVTZ level. Only a partial optimization to the potential grid is used to keep the atomic moments close to their DMA-derived values while still providing an improved molecular potential. The fitting approach is also useful for molecules containing symmetry-averaged atoms of the same atomic multipole type. In this case, simple arithmetic averaging would degrade the quality of the PAM. For example, in dimethyl- or trimethylamine, all the methyl hydrogen atoms are indistinguishable and adopt the same atom type. The DMA multipole values for these atoms are somewhat different due their non-equivalence in any single conformation, and PAM derived by simple averaging would lead to a large error in the molecular dipole moment. The potential-optimized PAM, where methyl hydrogens are constrained to adopt equivalent values, will reproduce almost exactly both the ab initio potential and the molecular multipole moments. Our standard procedure is to use a molecular potential grid consisting of a 2 Å shell beginning 1 Å out from the vdW surface. The DMA monopole values are generally fixed during the potential fitting procedure.
This electrostatic parameterization protocol is particularly important for larger molecules and for molecules with high symmetry. It is known that the original DMA approach tends to give “unphysical’ multipole values for large molecules when diffuse functions are included in the basis set even though the resulting electrostatic potential is correct. A recent modification of DMA80 has been put forward to address this issue. However, in our hands, the multipoles from the modified scheme seem less transferable between conformations. The above protocol allows derivation of PAM corresponding to larger basis sets than would be practical with the original DMA method. Note this procedure is different from restrained potential fits commonly used to fit fixed atomic charge models, as the starting DMA values are already quite reasonable and the fitting can be considered as a small perturbation biased toward the larger basis set potential. The overall procedure has been extensively tested in a small molecule hydration study81 (link) and will be used in future AMOEBA parameterization efforts.
Empirical vdW parameters were determined by fitting to both gas and liquid phase properties. The gas phase properties include homodimer binding energy (BSSE corrected) and structure from ab initio calculations at the MP2/aug-cc-pVTZ level or above. Liquid properties include experimental density and heat of vaporization of neat liquids. The vdW parameters were first estimated by comparing structure and energy of the AMOEBA-optimized dimer with ab initio results, and then fine-tuned to reproduce the experimental liquid density and heat of vaporization via molecular dynamics simulation. Additional homodimers at alternative configurations, heterodimers with water, and liquid properties were computed post facto for the purpose of validation. A more generic force field atom classification for vdW parameters was enforced to ensure the transferability. Table 1 lists the common vdW atom classes used by AMOEBA, together with the corresponding vdW parameters and polarizabilities. The vdW atom classes are also used to define parameters for all of the valence potential energy terms. The parameters for bonded terms, initially transferred from MM3, are optimized to reproduce ab initio geometries and vibration frequencies. In the final parameterization step, after all other parameters are fixed, torsional parameters were obtained by fitting to ab initio conformational energy profiles at the MP2/6-311++(2d,2p) level of theory.
Publication 2011

Most recents protocols related to «Methylamine»

Oligonucleotides
(5 μM) were incubated with 20 mM sodium phosphate (pH 6.0) and
20 mM DNFB in 40% DMSO for 30 min at 50 °C to activate 4sU. Subsequently,
the chloroform extraction was repeated seven times to eliminate residual
DNFB. The activated oligonucleotides were purified by HPLC (260 nm,
column: hydrosphere C18 250 × 4.6 mm2 (YMC), column
temperature: 50 °C, gradient: 0–40% solvent B for 15 min,
solvent A: 50 mM TEAA, 5% acetonitrile (pH 7.0), and solvent B: acetonitrile).
The speed-vac evaporator concentrated the activated oligonucleotides
and removed acetonitrile. To introduce amine groups, 20 mM sodium
borate (pH 8.5) and 180 mM methylamine or ammonium chloride were added
to the oligonucleotides, which were then incubated for 30 min at 50
°C. The amine-introduced oligonucleotides were prepared using
an HPLC system (260 nm, column: hydrosphere C18 250 × 4.6 mm2 (YMC), column temperature: 50 °C, gradient: 0–40%
solvent B for 15 min, solvent A: 50 mM TEAA, 5% acetonitrile (pH 7.0),
and solvent B: acetonitrile). UV spectra were obtained from the HPLC
results. Activated and base-converted oligonucleotides were analyzed
by using ESI-MS and MALDI-TOF-MS, respectively. The transcribed RNA
(50 nM RNA) was incubated with 20 mM sodium borate (pH 8.5) and 20
mM DNFB in 40% DMSO for 30 min at 50 °C to activate 4sU. Subsequently,
the chloroform extraction was repeated seven times to eliminate residual
DNFB. To introduce an amine group, 180 mM methylamine and 20 mM sodium
borate (pH 8.5) were added to the activated RNA and incubated for
30 min at 50 °C. Base-converted RNA was purified using alcohol
precipitation. The transcribed RNA after base activation and conversion
was analyzed by denaturing PAGE (7.5 M urea, 5% PAGE (acrylamide/N,N′-methylenebis(acrylamide) =
19:1)).
Publication 2024
4sUTP (5 μM) or transcribed
RNA (50 nM) was treated with DNFB and methylamine following the protocol
described above. The base-converted materials were digested at 37
°C for 16 h in 10 mM Tris–HCl (pH 8.0), 50 mM NaCl, 10
mM MgCl2, 1.2 mU/μL phosphodiesterase 1 (Sigma-Aldrich),
and 0.6 U/μL alkaline phosphatase (Calf intestine) (TaKaRa).
Then, the enzymes were eliminated by collecting flow-through from
an Amicon Ultra 10 K filter (MERCK). The degradants were analyzed
by HPLC (260 nm, column: hydrosphere C18 250 × 4.6 mm2 (YMC), column temperature: 50 °C, gradient: 10% solvent B for
5 min, 10–35% solvent B for 35 min, 80% solvent B for 5 min,
0% solvent B for 15 min, solvent A: 50 mM TEAA (pH 7.0), and solvent
B: 50 mM TEAA (pH 7.0), 50% acetonitrile).
Publication 2024
Methylamine (CH3NH2, Sigma Aldrich, ≥99.5%, St. Louis, MO, USA), silicon (Si, p-type, 10 mm × 10 mm, Siltron Inc., Seoul, Republic of Korea), tungsten (W) wire (thickness ~0.5 mm, The Nilaco Corporation, Tokyo, Japan), sodium dihydrogen phosphate (NaH2PO4, Sigma Aldrich, ≥99%, St. Louis, MO, USA), and disodium hydrogen phosphate (Na2HPO4, Sigma Aldrich, ≥99%, St. Louis, MO, USA) were used.
Publication 2024
To detect the presence of the hazardous methylamine, a three-electrode system of 10 mL electrochemical cells was employed for the measurement of cyclic voltammetry and linear sweep voltammetry using an electrometer (Keithley, 6517A, Aurora Rd, USA), Herein, HFCVD-grown WO3 NW thin film served as the working electrode, AgCl/Ag was employed as a reference electrode and gold wire was utilized as the counter electrode. A targeted analyte (methylamine) was prepared at a broad concentration range of 20 μM–1 mM in a 10 mL solution of 0.1 M phosphate buffer solution (PBS) of pH 7. The use of PBS in electrochemical sensing offers advantages such as stable pH, biocompatibility, consistent ionic strength, enhanced solubility, and compatibility. Its buffering capacity and versatility make PBS a reliable choice for maintaining optimal conditions during electrochemical measurements [37 (link)]. Cyclic voltammetry was performed to study the oxidation and reduction peaks and the linear sweep voltammetry technique was used to study the current and voltage responses. The active area of the fabricated electrode was 1 cm2 (WO3 NWs) and the sensitivity was calculated by dividing the slope of the calibrated current–concentration plot by the active area of the sensor, as expressed in Equation (1): Sensitivity=SlopeofcalibratedcurveActivearea
Herein, CV was performed within the range of −0.8~0.8 V, with a scan rate of 50 mV/s. The current responses were analyzed within a voltage range from 0 to 2.0 V, and the response time was determined to be 10 s.
Publication 2024
The perovskite solution was prepared according to our previous work29 (link),44 (link),45 (link). Take 3 mg mL−1 heptanal MAPbI3 perovskite solution as an example: at the beginning, 6 mg heptanal was added to 1 mL methylamine solution (MA, 33 wt.% in absolute ethanol) to obtain the 6 mg mL−1 heptanal-MA solution. Then 500 μL of the 6 mg mL−1 heptanal solution was added to a vial containing 159 mg MAI and 461 mg PbI2 powders. The vial was directly sonicated for 1 h until the solution became completely clear. It should be noted the stocked methylamine solution (MA, 33 wt.% in absolute ethanol) may have a lower concentration than 33 wt.% due to the evaporation of MA from ethanol. It is highly recommended to use fresh and new methylamine solution. After obtaining a clear solution from the abovementioned sonication, 500 μL ACN solvent was added to dilute the system. After a quick sonication for 5 min, the final 3 mg mL−1 heptanal MAPbI3 perovskite solution was ready to use within 1 day. Perovskite solutions with other biomolecules were prepared with the same process.
Publication 2024

Top products related to «Methylamine»

Sourced in United States, Germany
Methylamine is a colorless, flammable gas that is used as a chemical building block in the production of various pharmaceuticals and other industrial compounds. It serves as a key starting material in the synthesis of various medications and other chemical products.
Sourced in United States, Germany, United Kingdom, China, Italy, Sao Tome and Principe, France, Macao, India, Canada, Switzerland, Japan, Australia, Spain, Poland, Belgium, Brazil, Czechia, Portugal, Austria, Denmark, Israel, Sweden, Ireland, Hungary, Mexico, Netherlands, Singapore, Indonesia, Slovakia, Cameroon, Norway, Thailand, Chile, Finland, Malaysia, Latvia, New Zealand, Hong Kong, Pakistan, Uruguay, Bangladesh
DMSO is a versatile organic solvent commonly used in laboratory settings. It has a high boiling point, low viscosity, and the ability to dissolve a wide range of polar and non-polar compounds. DMSO's core function is as a solvent, allowing for the effective dissolution and handling of various chemical substances during research and experimentation.
Sourced in United States, Germany, United Kingdom, China, Italy, Canada, India, Australia, Japan, Spain, Singapore, Sao Tome and Principe, Poland, France, Switzerland, Macao, Chile, Belgium, Czechia, Hungary, Netherlands
N,N-dimethylformamide is a clear, colorless liquid organic compound with the chemical formula (CH3)2NC(O)H. It is a common laboratory solvent used in various chemical reactions and processes.
Sourced in Germany, United States, Italy, India, China, United Kingdom, France, Poland, Spain, Switzerland, Australia, Canada, Brazil, Sao Tome and Principe, Ireland, Belgium, Macao, Japan, Singapore, Mexico, Austria, Czechia, Bulgaria, Hungary, Egypt, Denmark, Chile, Malaysia, Israel, Croatia, Portugal, New Zealand, Romania, Norway, Sweden, Indonesia
Acetonitrile is a colorless, volatile, flammable liquid. It is a commonly used solvent in various analytical and chemical applications, including liquid chromatography, gas chromatography, and other laboratory procedures. Acetonitrile is known for its high polarity and ability to dissolve a wide range of organic compounds.
Sourced in United States
Hydroiodic acid is a clear, colorless, and fuming liquid chemical compound. It is an aqueous solution of hydrogen iodide. Hydroiodic acid is widely used in various industrial and laboratory applications.
Sourced in United States
Methylamine hydrochloride is a chemical compound used in various laboratory applications. It is a white, crystalline solid that is soluble in water and other polar solvents. Methylamine hydrochloride is commonly used as a reagent or an intermediate in organic synthesis and chemical research.
Sourced in Germany
Methylamine solution is a chemical product used in various industrial and laboratory applications. It is a clear, colorless liquid with a characteristic amine odor. Methylamine solution is a source of the methylamine chemical compound, which is a versatile intermediate in the synthesis of other organic compounds. The specific concentration and formulation of the methylamine solution may vary depending on the intended application.
Sourced in United States, Germany, China, Australia, Japan, Switzerland
Chlorobenzene is a colorless, volatile liquid used as an intermediate in the production of various chemicals. It serves as a precursor for the synthesis of other organic compounds.
Sourced in United States
Nano-W is a precision laboratory instrument designed for the observation and analysis of nanoscale materials and structures. It features advanced imaging capabilities, enabling high-resolution visualization of samples at the nanometer scale. The core function of Nano-W is to provide researchers with a powerful tool for exploring the properties and behavior of nanomaterials, supporting a wide range of scientific applications.
Sourced in United States, Germany, Italy, United Kingdom, India, Spain, Japan, Poland, France, Switzerland, Belgium, Canada, Portugal, China, Sweden, Singapore, Indonesia, Australia, Mexico, Brazil, Czechia
Toluene is a colorless, flammable liquid with a distinctive aromatic odor. It is a common organic solvent used in various industrial and laboratory applications. Toluene has a chemical formula of C6H5CH3 and is derived from the distillation of petroleum.

More about "Methylamine"

Methylaminea, DMSO, N,N-dimethylformamide, Acetonitrile, Hydroiodic acid, Methylamine hydrochloride, Methylamine solution, Chlorobenzene, Nano-W, Toluene, Methylation, Dimethylamine, Trimethylamine, Methylated derivatives, Pharmaceuticals, Agrochemicals, Industrial chemicals, Experimental protocols, Literature, Preprints, Patents, Reproducibility, Accuracy, Research optimization, Innovation