fbpx
Wikipedia

Hydrophobicity scales

Hydrophobicity scales are values that define the relative hydrophobicity or hydrophilicity of amino acid residues. The more positive the value, the more hydrophobic are the amino acids located in that region of the protein. These scales are commonly used to predict the transmembrane alpha-helices of membrane proteins. When consecutively measuring amino acids of a protein, changes in value indicate attraction of specific protein regions towards the hydrophobic region inside lipid bilayer.

The hydrophobic or hydrophilic character of a compound or amino acid is its hydropathic character,[1] hydropathicity, or hydropathy.

Hydrophobicity and the hydrophobic effect

 
Hydrogen bonds between molecules of liquid water

The hydrophobic effect represents the tendency of water to exclude non-polar molecules. The effect originates from the disruption of highly dynamic hydrogen bonds between molecules of liquid water. Polar chemical groups, such as OH group in methanol do not cause the hydrophobic effect. However, a pure hydrocarbon molecule, for example hexane, cannot accept or donate hydrogen bonds to water. Introduction of hexane into water causes disruption of the hydrogen bonding network between water molecules. The hydrogen bonds are partially reconstructed by building a water "cage" around the hexane molecule, similar to that in clathrate hydrates formed at lower temperatures. The mobility of water molecules in the "cage" (or solvation shell) is strongly restricted. This leads to significant losses in translational and rotational entropy of water molecules and makes the process unfavorable in terms of free energy of the system.[2][3][4][5] In terms of thermodynamics, the hydrophobic effect is the free energy change of water surrounding a solute.[6] A positive free energy change of the surrounding solvent indicates hydrophobicity, whereas a negative free energy change implies hydrophilicity. In this way, the hydrophobic effect not only can be localized but also decomposed into enthalpic and entropic contributions.

Types of amino acid hydrophobicity scales

 
A table comparing four different scales for the hydrophobicity of an amino acid residue in a protein with the most hydrophobic amino acids on the top

A number of different hydrophobicity scales have been developed.[3][1][7][8][9]

There are clear differences between the four scales shown in the table.[10] Both the second and fourth scales place cysteine as the most hydrophobic residue, unlike the other two scales. This difference is due to the different methods used to measure hydrophobicity. The method used to obtain the Janin and Rose et al. scales was to examine proteins with known 3-D structures and define the hydrophobic character as the tendency for a residue to be found inside of a protein rather than on its surface.[11][12] Since cysteine forms disulfide bonds that must occur inside a globular structure, cysteine is ranked as the most hydrophobic. The first and third scales are derived from the physiochemical properties of the amino acid side chains. These scales result mainly from inspection of the amino acid structures.[13][1] Biswas et al., divided the scales based on the method used to obtain the scale into five different categories.[3]

Partitioning methods

The most common method of measuring amino acid hydrophobicity is partitioning between two immiscible liquid phases. Different organic solvents are most widely used to mimic the protein interior. However, organic solvents are slightly miscible with water and the characteristics of both phases change making it difficult to obtain pure hydrophobicity scale.[3] Nozaki and Tanford proposed the first major hydrophobicity scale for nine amino acids.[14] Ethanol and dioxane are used as the organic solvents and the free energy of transfer of each amino acid was calculated. Non liquid phases can also be used with partitioning methods such as micellar phases and vapor phases. Two scales have been developed using micellar phases.[15][16] Fendler et al. measured the partitioning of 14 radiolabeled amino acids using sodium dodecyl sulfate (SDS) micelles. Also, amino acid side chain affinity for water was measured using vapor phases.[13] Vapor phases represent the simplest non polar phases, because it has no interaction with the solute.[17] The hydration potential and its correlation to the appearance of amino acids on the surface of proteins was studied by Wolfenden. Aqueous and polymer phases were used in the development of a novel partitioning scale.[18] Partitioning methods have many drawbacks. First, it is difficult to mimic the protein interior.[19][20] In addition, the role of self solvation makes using free amino acids very difficult. Moreover, hydrogen bonds that are lost in the transfer to organic solvents are not reformed but often in the interior of protein.[21]

Accessible surface area methods

Hydrophobicity scales can also be obtained by calculating the solvent accessible surface areas for amino acid residues in the expended polypeptide chain[21] or in alpha-helix and multiplying the surface areas by the empirical solvation parameters for the corresponding types of atoms.[3] A differential solvent accessible surface area hydrophobicity scale based on proteins as compacted networks near a critical point, due to self-organization by evolution, was constructed based on asymptotic power-law (self-similar) behavior.[22][23] This scale is based on a bioinformatic survey of 5526 high-resolution structures from the Protein Data Bank. This differential scale has two comparative advantages: (1) it is especially useful for treating changes in water-protein interactions that are too small to be accessible to conventional force-field calculations, and (2) for homologous structures, it can yield correlations with changes in properties from mutations in the amino acid sequences alone, without determining corresponding structural changes, either in vitro or in vivo.

Chromatographic methods

Reversed phase liquid chromatography (RPLC) is the most important chromatographic method for measuring solute hydrophobicity.[3][24] The non polar stationary phase mimics biological membranes. Peptide usage has many advantages because partition is not extended by the terminal charges in RPLC. Also, secondary structures formation is avoided by using short sequence peptides. Derivatization of amino acids is necessary to ease its partition into a C18 bonded phase. Another scale had been developed in 1971 and used peptide retention on hydrophilic gel.[25] 1-butanol and pyridine were used as the mobile phase in this particular scale and glycine was used as the reference value. Pliska and his coworkers[26] used thin layer chromatography to relate mobility values of free amino acids to their hydrophobicities. About a decade ago, another hydrophilicity scale was published, this scale used normal phase liquid chromatography and showed the retention of 121 peptides on an amide-80 column.[27] The absolute values and relative rankings of hydrophobicity determined by chromatographic methods can be affected by a number of parameters. These parameters include the silica surface area and pore diameter, the choice and pH of aqueous buffer, temperature and the bonding density of stationary phase chains.[3] ip mw hydrophobicity proteins

Site-directed mutagenesis

This method use DNA recombinant technology and it gives an actual measurement of protein stability. In his detailed site-directed mutagenesis studies, Utani and his coworkers substituted 19 amino acids at Trp49 of the tryptophan synthase and he measured the free energy of unfolding. They found that the increased stability is directly proportional to increase in hydrophobicity up to a certain size limit. The main disadvantage of site-directed mutagenesis method is that not all the 20 naturally occurring amino acids can substitute a single residue in a protein. Moreover, these methods have cost problems and is useful only for measuring protein stability.[3][28]

Physical property methods

 
Wimley-White whole-residue hydrophobicity scales

The hydrophobicity scales developed by physical property methods are based on the measurement of different physical properties. Examples include, partial molar heat capacity, transition temperature and surface tension. Physical methods are easy to use and flexible in terms of solute. The most popular hydrophobicity scale was developed by measuring surface tension values for the naturally occurring 20 amino acids in NaCl solution.[29] The main drawbacks of surface tension measurements is that the broken hydrogen bonds and the neutralized charged groups remain at the solution air interface.[3][1] Another physical property method involve measuring the solvation free energy.[30] The solvation free energy is estimated as a product of an accessibility of an atom to the solvent and an atomic solvation parameter. Results indicate the solvation free energy lowers by an average of 1 Kcal/residue upon folding.[3]

 
Whole-residue octanol-scale hydropathy plot for the L-subunit of the photosynthetic reaction center of Rhodobacter sphaeroides.

Recent applications

Palliser and Parry have examined about 100 scales and found that they can use them for locating B-strands on the surface of proteins.[31] Hydrophobicity scales were also used to predict the preservation of the genetic code.[32] Trinquier observed a new order of the bases that better reflect the conserved character of the genetic code.[3] They believed new ordering of the bases was uracil-guanine-cystosine-adenine(UGCA)better reflected the conserved character of the genetic code compared to the commonly seen ordering UCAG.[3]

Wimley–White whole residue hydrophobicity scales

The Wimley–White whole residue hydrophobicity scales are significant for two reasons. First, they include the contributions of the peptide bonds as well as the sidechains, providing absolute values. Second, they are based on direct, experimentally determined values for transfer free energies of polypeptides.

Two whole-residue hydrophobicity scales have been measured:

  • One for the transfer of unfolded chains from water to the bilayer interface (referred to as the Wimley–White interfacial hydrophobicity scale).
  • One for the transfer of unfolded chains into octanol, which is relevant to the hydrocarbon core of a bilayer.

The Stephen H. White website[33] provides an example of whole residue hydrophobicity scales showing the free energy of transfer ΔG(kcal/mol) from water to POPC interface and to n-octanol.[33] These two scales are then used together to make Whole residue hydropathy plots.[33] The hydropathy plot constructed using ΔGwoct − ΔGwif shows favorable peaks on the absolute scale that correspond to the known TM helices. Thus, the whole residue hydropathy plots illustrate why transmembrane segments prefer a transmembrane location rather than a surface one.[34][35][36][37]

Amino acid Interface scale,
ΔGwif (kcal/mol)
Octanol scale,
ΔGwoct (kcal/mol)
Octanol − interface,
ΔGwoct − ΔGwif
Ile −0.31 −1.12 −0.81
Leu −0.56 −1.25 −0.69
Phe −1.13 −1.71 −0.58
Val 0.07 −0.46 −0.53
Met −0.23 −0.67 −0.44
Pro 0.45 0.14 −0.31
Trp −1.85 −2.09 −0.24
His0 0.17 0.11 −0.06
Thr 0.14 0.25 0.11
Glu0 −0.01 0.11 0.12
Gln 0.58 0.77 0.19
Cys −0.24 −0.02 0.22
Tyr −0.94 −0.71 0.23
Ala 0.17 0.50 0.33
Ser 0.13 0.46 0.33
Asn 0.42 0.85 0.43
Asp0 −0.07 0.43 0.50
Arg+ 0.81 1.81 1.00
Gly 0.01 1.15 1.14
His+ 0.96 2.33 1.37
Glu- 2.02 3.63 1.61
Lys+ 0.99 2.80 1.81
Asp- 1.23 3.64 2.41

Bandyopadhyay-Mehler protein structure based scales

Most of the existing hydrophobicity scales are derived from the properties of amino acids in their free forms or as a part of a short peptide. Bandyopadhyay-Mehler hydrophobicity scale was based on partitioning of amino acids in the context of protein structure. Protein structure is a complex mosaic of various dielectric medium generated by arrangement of different amino acids. Hence, different parts of the protein structure most likely would behave as solvents with different dielectric values. For simplicity, each protein structure was considered as an immiscible mixture of two solvents, protein interior and protein exterior. The local environment around individual amino acid (termed as "micro-environment") was computed for both protein interior and protein exterior. The ratio gives the relative hydrophobicity scale for individual amino acids. Computation was trained on high resolution protein crystal structures. This quantitative descriptor for microenvironment was derived from the octanol-water partition coefficient, (known as Rekker's Fragmental Constants) widely used for pharmacophores. This scale well correlate with the existing methods, based on partitioning and free energy computations. Advantage of this scale is it is more realistic, as it is in the context of real protein structures.[9]

Scale based on contact angle of water nanodroplet

 
Contact angles of a water nanodroplet on the artificial beta-sheets with various amino acid side chains
 
The MD simulation system and the structure of artificial beta-folding 2D peptide network composed of unified R-side chains.

In the field of engineering, the hydrophobicity (or dewetting ability) of a flat surface (e.g., a counter top in kitchen or a cooking pan) can be measured by the contact angle of water droplet. A University of Nebraska-Lincoln team recently devised a computational approach that can relate the molecular hydrophobicity scale of amino-acid chains to the contact angle of water nanodroplet.[38] The team constructed planar networks composed of unified amino-acid side chains with native structure of the beta-sheet protein. Using molecular dynamics simulation, the team is able to measure the contact angle of water nanodroplet on the planar networks (caHydrophobicity).

On the other hand, previous studies show that the minimum of excess chemical potential of a hard-sphere solute with respect to that in the bulk exhibits a linear dependence on cosine value of contact angle.[39] Based on the computed excess chemical potentials of the purely repulsive methane-sized Weeks–Chandler–Andersen solute with respect to that in the bulk, the extrapolated values of cosine value of contact angle are calculated(ccHydrophobicity), which can be used to quantify the hydrophobicity of amino acid side chains with complete wetting behaviors.

See also

References

  1. ^ a b c d Kyte, Jack; Doolittle, Russell F. (May 1982). "A simple method for displaying the hydropathic character of a protein". Journal of Molecular Biology. Elsevier BV. 157 (1): 105–32. CiteSeerX 10.1.1.458.454. doi:10.1016/0022-2836(82)90515-0. PMID 7108955.
  2. ^ Tanford, C., The hydrophobic effect(New York:Wiley.1980).
  3. ^ a b c d e f g h i j k l Biswas, Kallol M.; DeVido, Daniel R.; Dorsey, John G. (2003). "Evaluation of methods for measuring amino acid hydrophobicities and interactions". Journal of Chromatography A. Elsevier BV. 1000 (1–2): 637–655. doi:10.1016/s0021-9673(03)00182-1. ISSN 0021-9673. PMID 12877193.
  4. ^ W . Kauzmann, Adv. Protein Chem. 14 (1959) 1.
  5. ^ Charton, Marvin; Charton, Barbara I. (1982). "The structural dependence of amino acid hydrophobicity parameters". Journal of Theoretical Biology. Elsevier BV. 99 (4): 629–644. Bibcode:1982JThBi..99..629C. doi:10.1016/0022-5193(82)90191-6. ISSN 0022-5193. PMID 7183857.
  6. ^ Schauperl, M; Podewitz, M; Waldner, BJ; Liedl, KR (2016). "Enthalpic and Entropic Contributions to Hydrophobicity". Journal of Chemical Theory and Computation. 12 (9): 4600–10. doi:10.1021/acs.jctc.6b00422. PMC 5024328. PMID 27442443.
  7. ^ Eisenberg D (July 1984). "Three-dimensional structure of membrane and surface proteins". Annu. Rev. Biochem. 53: 595–623. doi:10.1146/annurev.bi.53.070184.003115. PMID 6383201.
  8. ^ Rose, G D; Wolfenden, R (1993). "Hydrogen Bonding, Hydrophobicity, Packing, and Protein Folding". Annual Review of Biophysics and Biomolecular Structure. Annual Reviews. 22 (1): 381–415. doi:10.1146/annurev.bb.22.060193.002121. ISSN 1056-8700. PMID 8347995.
  9. ^ a b Bandyopadhyay, D., Mehler, E.L. (2008). "Quantitative expression of protein heterogeneity: Response of amino acid side chains to their local environment". Proteins: Structure, Function, and Bioinformatics. 72 (2): 646–659. doi:10.1002/prot.21958. PMID 18247345. S2CID 20929779.
  10. ^ "Hydrophobicity Scales".
  11. ^ Janin, Joël (1979). "Surface and inside volumes in globular proteins". Nature. Springer Science and Business Media LLC. 277 (5696): 491–492. Bibcode:1979Natur.277..491J. doi:10.1038/277491a0. ISSN 0028-0836. PMID 763335. S2CID 4338901.
  12. ^ Rose, G.; Geselowitz, A.; Lesser, G.; Lee, R.; Zehfus, M. (1985-08-30). "Hydrophobicity of amino acid residues in globular proteins". Science. American Association for the Advancement of Science (AAAS). 229 (4716): 834–838. Bibcode:1985Sci...229..834R. doi:10.1126/science.4023714. ISSN 0036-8075. PMID 4023714.
  13. ^ a b Wolfenden, R.; Andersson, L.; Cullis, P. M.; Southgate, C. C. B. (1981). "Affinities of amino acid side chains for solvent water". Biochemistry. American Chemical Society (ACS). 20 (4): 849–855. doi:10.1021/bi00507a030. ISSN 0006-2960. PMID 7213619.
  14. ^ Y . Nozaki, C. Tanford, J. Biol. Chem. 246 (1971) 2211.
  15. ^ Fendler, Janos H.; Nome, Faruk; Nagyvary, Joseph (1975). "Compartmentalization of amino acids in surfactant aggregates". Journal of Molecular Evolution. Springer Science and Business Media LLC. 6 (3): 215–232. Bibcode:1975JMolE...6..215F. doi:10.1007/bf01732358. ISSN 0022-2844. PMID 1206727. S2CID 2394979.
  16. ^ Leodidis, Epaminondas B.; Hatton, T. Alan. (1990). "Amino acids in AOT reversed micelles. 2. The hydrophobic effect and hydrogen bonding as driving forces for interfacial solubilization". The Journal of Physical Chemistry. American Chemical Society (ACS). 94 (16): 6411–6420. doi:10.1021/j100379a047. ISSN 0022-3654.
  17. ^ Sharp, Kim A.; Nicholls, Anthony; Friedman, Richard; Honig, Barry (1991-10-08). "Extracting hydrophobic free energies from experimental data: relationship to protein folding and theoretical models". Biochemistry. American Chemical Society (ACS). 30 (40): 9686–9697. doi:10.1021/bi00104a017. ISSN 0006-2960. PMID 1911756.
  18. ^ Zaslavsky, B. Yu.; Mestechkina, N.M.; Miheeva, L.M.; Rogozhin, S.V. (1982). "Measurement of relative hydrophobicity of amino acid side-chains by partition in an aqueous two-phase polymeric system: Hydrophobicity scale for non-polar and ionogenic side-chains". Journal of Chromatography A. Elsevier BV. 240 (1): 21–28. doi:10.1016/s0021-9673(01)84003-6. ISSN 0021-9673.
  19. ^ S . Damadoran, K.B. Song, J. Biol. Chem. 261 (1986) 7220.
  20. ^ Ben-Naim, A. (1990-02-15). "Solvent effects on protein association and protein folding". Biopolymers. Wiley. 29 (3): 567–596. doi:10.1002/bip.360290312. ISSN 0006-3525. PMID 2331515. S2CID 25691137.
  21. ^ a b Chothia, Cyrus (1976). "The nature of the accessible and buried surfaces in proteins". Journal of Molecular Biology. Elsevier BV. 105 (1): 1–12. doi:10.1016/0022-2836(76)90191-1. ISSN 0022-2836. PMID 994183.
  22. ^ Moret, M. A.; Zebende, G. F. (2007-01-19). "Amino acid hydrophobicity and accessible surface area". Physical Review E. American Physical Society (APS). 75 (1): 011920. Bibcode:2007PhRvE..75a1920M. doi:10.1103/physreve.75.011920. ISSN 1539-3755. PMID 17358197.
  23. ^ Phillips, J. C. (2009-11-20). "Scaling and self-organized criticality in proteins: Lysozymec". Physical Review E. American Physical Society (APS). 80 (5): 051916. Bibcode:2009PhRvE..80e1916P. doi:10.1103/physreve.80.051916. ISSN 1539-3755. PMID 20365015.
  24. ^ Hodges, Robert S.; Zhu, Bing-Yan; Zhou, Nian E.; Mant, Colin T. (1994). "Reversed-phase liquid chromatography as a useful probe of hydrophobic interactions involved in protein folding and protein stability". Journal of Chromatography A. Elsevier BV. 676 (1): 3–15. doi:10.1016/0021-9673(94)80452-4. ISSN 0021-9673. PMID 7921179.
  25. ^ Aboderin, Akintola A. (1971). "An empirical hydrophobicity scale for α-amino-acids and some of its applications". International Journal of Biochemistry. Elsevier BV. 2 (11): 537–544. doi:10.1016/0020-711x(71)90023-1. ISSN 0020-711X.
  26. ^ Pliška, Vladimir; Schmidt, Manfred; Fauchère, Jean-Luc (1981). "Partition coefficients of amino acids and hydrophobic parameters π of their side-chains as measured by thin-layer chromatography". Journal of Chromatography A. Elsevier BV. 216: 79–92. doi:10.1016/s0021-9673(00)82337-7. ISSN 0021-9673.
  27. ^ Plass, Monika; Valko, Klara; Abraham, Michael H (1998). "Determination of solute descriptors of tripeptide derivatives based on high-throughput gradient high-performance liquid chromatography retention data". Journal of Chromatography A. Elsevier BV. 803 (1–2): 51–60. doi:10.1016/s0021-9673(97)01215-6. ISSN 0021-9673.
  28. ^ Yutani, K.; Ogasahara, K.; Tsujita, T.; Sugino, Y. (1987-07-01). "Dependence of conformational stability on hydrophobicity of the amino acid residue in a series of variant proteins substituted at a unique position of tryptophan synthase alpha subunit". Proceedings of the National Academy of Sciences USA. Proceedings of the National Academy of Sciences. 84 (13): 4441–4444. Bibcode:1987PNAS...84.4441Y. doi:10.1073/pnas.84.13.4441. ISSN 0027-8424. PMC 305105. PMID 3299367.
  29. ^ Bull, Henry B.; Breese, Keith (1974). "Surface tension of amino acid solutions: A hydrophobicity scale of the amino acid residues". Archives of Biochemistry and Biophysics. Elsevier BV. 161 (2): 665–670. doi:10.1016/0003-9861(74)90352-x. ISSN 0003-9861. PMID 4839053.
  30. ^ Eisenberg, David; McLachlan, Andrew D. (1986). "Solvation energy in protein folding and binding". Nature. Springer Science and Business Media LLC. 319 (6050): 199–203. Bibcode:1986Natur.319..199E. doi:10.1038/319199a0. ISSN 0028-0836. PMID 3945310. S2CID 21867582.
  31. ^ Palliser, Christopher C.; Parry, David A. D. (2000). "Quantitative comparison of the ability of hydropathy scales to recognize surface ?-strands in proteins". Proteins: Structure, Function, and Genetics. Wiley. 42 (2): 243–255. doi:10.1002/1097-0134(20010201)42:2<243::aid-prot120>3.0.co;2-b. ISSN 0887-3585. PMID 11119649. S2CID 23839522.
  32. ^ G . Trinquier, Y.-H. Sanejouand, Protein Eng. 11 (1998) 153.
  33. ^ a b c White, Stephen (2006-06-29). "Experimentally Determined Hydrophobicity Scales". University of California, Irvine. Retrieved 2009-06-12.
  34. ^ Wimley, William C.; White, Stephen H. (1996). "Experimentally determined hydrophobicity scale for proteins at membrane interfaces". Nature Structural & Molecular Biology. Springer Science and Business Media LLC. 3 (10): 842–848. doi:10.1038/nsb1096-842. ISSN 1545-9993. PMID 8836100. S2CID 1823375.
  35. ^ Wimley, William C.; Creamer, Trevor P.; White, Stephen H. (1996). "Solvation Energies of Amino Acid Side Chains and Backbone in a Family of Host−Guest Pentapeptides". Biochemistry. American Chemical Society (ACS). 35 (16): 5109–5124. doi:10.1021/bi9600153. ISSN 0006-2960. PMID 8611495.
  36. ^ White SH. & Wimley WC (1998). Biochim. Biophys. Acta 1376:339-352.
  37. ^ White, Stephen H.; Wimley, William C. (1999). "MEMBRANE PROTEIN FOLDING AND STABILITY: Physical Principles". Annual Review of Biophysics and Biomolecular Structure. Annual Reviews. 28 (1): 319–365. doi:10.1146/annurev.biophys.28.1.319. ISSN 1056-8700. PMID 10410805.
  38. ^ Zhu, Chongqin; Gao, Yurui; Li, Hui; Meng, Sheng; Li, Lei; Francisco, Joseph S; Zeng, Xiao Cheng (2016). "Characterizing hydrophobicity of amino acid side chains in a protein environment via measuring contact angle of a water nanodroplet on planar peptide network". Proceedings of the National Academy of Sciences. 113 (46): 12946–12951. doi:10.1073/pnas.1616138113. PMC 5135335. PMID 27803319.
  39. ^ Godawat, R; Jamadagni, S. N; Garde, S (2009). "Characterizing hydrophobicity of interfaces by using cavity formation, solute binding, and water correlations". Proceedings of the National Academy of Sciences. 106 (36): 15119–15124. doi:10.1073/pnas.0902778106. PMC 2741215. PMID 19706896.

External links

  • ProtScale (web-based tool for calculating hydropathy plots)
  • NetSurfP - Secondary Structure and Surface accessibility predictor
  • Whole residue hydrophobicity scale
  • Membrane protein explorer

hydrophobicity, scales, values, that, define, relative, hydrophobicity, hydrophilicity, amino, acid, residues, more, positive, value, more, hydrophobic, amino, acids, located, that, region, protein, these, scales, commonly, used, predict, transmembrane, alpha,. Hydrophobicity scales are values that define the relative hydrophobicity or hydrophilicity of amino acid residues The more positive the value the more hydrophobic are the amino acids located in that region of the protein These scales are commonly used to predict the transmembrane alpha helices of membrane proteins When consecutively measuring amino acids of a protein changes in value indicate attraction of specific protein regions towards the hydrophobic region inside lipid bilayer The hydrophobic or hydrophilic character of a compound or amino acid is its hydropathic character 1 hydropathicity or hydropathy Contents 1 Hydrophobicity and the hydrophobic effect 2 Types of amino acid hydrophobicity scales 3 Partitioning methods 4 Accessible surface area methods 5 Chromatographic methods 6 Site directed mutagenesis 7 Physical property methods 8 Recent applications 9 Wimley White whole residue hydrophobicity scales 10 Bandyopadhyay Mehler protein structure based scales 11 Scale based on contact angle of water nanodroplet 12 See also 13 References 14 External linksHydrophobicity and the hydrophobic effect EditMain article Hydrophobic effect Hydrogen bonds between molecules of liquid water The hydrophobic effect represents the tendency of water to exclude non polar molecules The effect originates from the disruption of highly dynamic hydrogen bonds between molecules of liquid water Polar chemical groups such as OH group in methanol do not cause the hydrophobic effect However a pure hydrocarbon molecule for example hexane cannot accept or donate hydrogen bonds to water Introduction of hexane into water causes disruption of the hydrogen bonding network between water molecules The hydrogen bonds are partially reconstructed by building a water cage around the hexane molecule similar to that in clathrate hydrates formed at lower temperatures The mobility of water molecules in the cage or solvation shell is strongly restricted This leads to significant losses in translational and rotational entropy of water molecules and makes the process unfavorable in terms of free energy of the system 2 3 4 5 In terms of thermodynamics the hydrophobic effect is the free energy change of water surrounding a solute 6 A positive free energy change of the surrounding solvent indicates hydrophobicity whereas a negative free energy change implies hydrophilicity In this way the hydrophobic effect not only can be localized but also decomposed into enthalpic and entropic contributions Types of amino acid hydrophobicity scales Edit A table comparing four different scales for the hydrophobicity of an amino acid residue in a protein with the most hydrophobic amino acids on the top A number of different hydrophobicity scales have been developed 3 1 7 8 9 There are clear differences between the four scales shown in the table 10 Both the second and fourth scales place cysteine as the most hydrophobic residue unlike the other two scales This difference is due to the different methods used to measure hydrophobicity The method used to obtain the Janin and Rose et al scales was to examine proteins with known 3 D structures and define the hydrophobic character as the tendency for a residue to be found inside of a protein rather than on its surface 11 12 Since cysteine forms disulfide bonds that must occur inside a globular structure cysteine is ranked as the most hydrophobic The first and third scales are derived from the physiochemical properties of the amino acid side chains These scales result mainly from inspection of the amino acid structures 13 1 Biswas et al divided the scales based on the method used to obtain the scale into five different categories 3 Partitioning methods EditThe most common method of measuring amino acid hydrophobicity is partitioning between two immiscible liquid phases Different organic solvents are most widely used to mimic the protein interior However organic solvents are slightly miscible with water and the characteristics of both phases change making it difficult to obtain pure hydrophobicity scale 3 Nozaki and Tanford proposed the first major hydrophobicity scale for nine amino acids 14 Ethanol and dioxane are used as the organic solvents and the free energy of transfer of each amino acid was calculated Non liquid phases can also be used with partitioning methods such as micellar phases and vapor phases Two scales have been developed using micellar phases 15 16 Fendler et al measured the partitioning of 14 radiolabeled amino acids using sodium dodecyl sulfate SDS micelles Also amino acid side chain affinity for water was measured using vapor phases 13 Vapor phases represent the simplest non polar phases because it has no interaction with the solute 17 The hydration potential and its correlation to the appearance of amino acids on the surface of proteins was studied by Wolfenden Aqueous and polymer phases were used in the development of a novel partitioning scale 18 Partitioning methods have many drawbacks First it is difficult to mimic the protein interior 19 20 In addition the role of self solvation makes using free amino acids very difficult Moreover hydrogen bonds that are lost in the transfer to organic solvents are not reformed but often in the interior of protein 21 Accessible surface area methods EditMain article Implicit solvation Hydrophobicity scales can also be obtained by calculating the solvent accessible surface areas for amino acid residues in the expended polypeptide chain 21 or in alpha helix and multiplying the surface areas by the empirical solvation parameters for the corresponding types of atoms 3 A differential solvent accessible surface area hydrophobicity scale based on proteins as compacted networks near a critical point due to self organization by evolution was constructed based on asymptotic power law self similar behavior 22 23 This scale is based on a bioinformatic survey of 5526 high resolution structures from the Protein Data Bank This differential scale has two comparative advantages 1 it is especially useful for treating changes in water protein interactions that are too small to be accessible to conventional force field calculations and 2 for homologous structures it can yield correlations with changes in properties from mutations in the amino acid sequences alone without determining corresponding structural changes either in vitro or in vivo Chromatographic methods EditReversed phase liquid chromatography RPLC is the most important chromatographic method for measuring solute hydrophobicity 3 24 The non polar stationary phase mimics biological membranes Peptide usage has many advantages because partition is not extended by the terminal charges in RPLC Also secondary structures formation is avoided by using short sequence peptides Derivatization of amino acids is necessary to ease its partition into a C18 bonded phase Another scale had been developed in 1971 and used peptide retention on hydrophilic gel 25 1 butanol and pyridine were used as the mobile phase in this particular scale and glycine was used as the reference value Pliska and his coworkers 26 used thin layer chromatography to relate mobility values of free amino acids to their hydrophobicities About a decade ago another hydrophilicity scale was published this scale used normal phase liquid chromatography and showed the retention of 121 peptides on an amide 80 column 27 The absolute values and relative rankings of hydrophobicity determined by chromatographic methods can be affected by a number of parameters These parameters include the silica surface area and pore diameter the choice and pH of aqueous buffer temperature and the bonding density of stationary phase chains 3 ip mw hydrophobicity proteinsSite directed mutagenesis EditThis method use DNA recombinant technology and it gives an actual measurement of protein stability In his detailed site directed mutagenesis studies Utani and his coworkers substituted 19 amino acids at Trp49 of the tryptophan synthase and he measured the free energy of unfolding They found that the increased stability is directly proportional to increase in hydrophobicity up to a certain size limit The main disadvantage of site directed mutagenesis method is that not all the 20 naturally occurring amino acids can substitute a single residue in a protein Moreover these methods have cost problems and is useful only for measuring protein stability 3 28 Physical property methods Edit Wimley White whole residue hydrophobicity scales The hydrophobicity scales developed by physical property methods are based on the measurement of different physical properties Examples include partial molar heat capacity transition temperature and surface tension Physical methods are easy to use and flexible in terms of solute The most popular hydrophobicity scale was developed by measuring surface tension values for the naturally occurring 20 amino acids in NaCl solution 29 The main drawbacks of surface tension measurements is that the broken hydrogen bonds and the neutralized charged groups remain at the solution air interface 3 1 Another physical property method involve measuring the solvation free energy 30 The solvation free energy is estimated as a product of an accessibility of an atom to the solvent and an atomic solvation parameter Results indicate the solvation free energy lowers by an average of 1 Kcal residue upon folding 3 Whole residue octanol scale hydropathy plot for the L subunit of the photosynthetic reaction center of Rhodobacter sphaeroides Recent applications EditPalliser and Parry have examined about 100 scales and found that they can use them for locating B strands on the surface of proteins 31 Hydrophobicity scales were also used to predict the preservation of the genetic code 32 Trinquier observed a new order of the bases that better reflect the conserved character of the genetic code 3 They believed new ordering of the bases was uracil guanine cystosine adenine UGCA better reflected the conserved character of the genetic code compared to the commonly seen ordering UCAG 3 Wimley White whole residue hydrophobicity scales EditThe Wimley White whole residue hydrophobicity scales are significant for two reasons First they include the contributions of the peptide bonds as well as the sidechains providing absolute values Second they are based on direct experimentally determined values for transfer free energies of polypeptides Two whole residue hydrophobicity scales have been measured One for the transfer of unfolded chains from water to the bilayer interface referred to as the Wimley White interfacial hydrophobicity scale One for the transfer of unfolded chains into octanol which is relevant to the hydrocarbon core of a bilayer The Stephen H White website 33 provides an example of whole residue hydrophobicity scales showing the free energy of transfer DG kcal mol from water to POPC interface and to n octanol 33 These two scales are then used together to make Whole residue hydropathy plots 33 The hydropathy plot constructed using DGwoct DGwif shows favorable peaks on the absolute scale that correspond to the known TM helices Thus the whole residue hydropathy plots illustrate why transmembrane segments prefer a transmembrane location rather than a surface one 34 35 36 37 Amino acid Interface scale DGwif kcal mol Octanol scale DGwoct kcal mol Octanol interface DGwoct DGwifIle 0 31 1 12 0 81Leu 0 56 1 25 0 69Phe 1 13 1 71 0 58Val 0 07 0 46 0 53Met 0 23 0 67 0 44Pro 0 45 0 14 0 31Trp 1 85 2 09 0 24His0 0 17 0 11 0 06Thr 0 14 0 25 0 11Glu0 0 01 0 11 0 12Gln 0 58 0 77 0 19Cys 0 24 0 02 0 22Tyr 0 94 0 71 0 23Ala 0 17 0 50 0 33Ser 0 13 0 46 0 33Asn 0 42 0 85 0 43Asp0 0 07 0 43 0 50Arg 0 81 1 81 1 00Gly 0 01 1 15 1 14His 0 96 2 33 1 37Glu 2 02 3 63 1 61Lys 0 99 2 80 1 81Asp 1 23 3 64 2 41Bandyopadhyay Mehler protein structure based scales EditMost of the existing hydrophobicity scales are derived from the properties of amino acids in their free forms or as a part of a short peptide Bandyopadhyay Mehler hydrophobicity scale was based on partitioning of amino acids in the context of protein structure Protein structure is a complex mosaic of various dielectric medium generated by arrangement of different amino acids Hence different parts of the protein structure most likely would behave as solvents with different dielectric values For simplicity each protein structure was considered as an immiscible mixture of two solvents protein interior and protein exterior The local environment around individual amino acid termed as micro environment was computed for both protein interior and protein exterior The ratio gives the relative hydrophobicity scale for individual amino acids Computation was trained on high resolution protein crystal structures This quantitative descriptor for microenvironment was derived from the octanol water partition coefficient known as Rekker s Fragmental Constants widely used for pharmacophores This scale well correlate with the existing methods based on partitioning and free energy computations Advantage of this scale is it is more realistic as it is in the context of real protein structures 9 Scale based on contact angle of water nanodroplet Edit Contact angles of a water nanodroplet on the artificial beta sheets with various amino acid side chains The MD simulation system and the structure of artificial beta folding 2D peptide network composed of unified R side chains In the field of engineering the hydrophobicity or dewetting ability of a flat surface e g a counter top in kitchen or a cooking pan can be measured by the contact angle of water droplet A University of Nebraska Lincoln team recently devised a computational approach that can relate the molecular hydrophobicity scale of amino acid chains to the contact angle of water nanodroplet 38 The team constructed planar networks composed of unified amino acid side chains with native structure of the beta sheet protein Using molecular dynamics simulation the team is able to measure the contact angle of water nanodroplet on the planar networks caHydrophobicity On the other hand previous studies show that the minimum of excess chemical potential of a hard sphere solute with respect to that in the bulk exhibits a linear dependence on cosine value of contact angle 39 Based on the computed excess chemical potentials of the purely repulsive methane sized Weeks Chandler Andersen solute with respect to that in the bulk the extrapolated values of cosine value of contact angle are calculated ccHydrophobicity which can be used to quantify the hydrophobicity of amino acid side chains with complete wetting behaviors See also EditHydrophobic mismatchReferences Edit a b c d Kyte Jack Doolittle Russell F May 1982 A simple method for displaying the hydropathic character of a protein Journal of Molecular Biology Elsevier BV 157 1 105 32 CiteSeerX 10 1 1 458 454 doi 10 1016 0022 2836 82 90515 0 PMID 7108955 Tanford C The hydrophobic effect New York Wiley 1980 a b c d e f g h i j k l Biswas Kallol M DeVido Daniel R Dorsey John G 2003 Evaluation of methods for measuring amino acid hydrophobicities and interactions Journal of Chromatography A Elsevier BV 1000 1 2 637 655 doi 10 1016 s0021 9673 03 00182 1 ISSN 0021 9673 PMID 12877193 W Kauzmann Adv Protein Chem 14 1959 1 Charton Marvin Charton Barbara I 1982 The structural dependence of amino acid hydrophobicity parameters Journal of Theoretical Biology Elsevier BV 99 4 629 644 Bibcode 1982JThBi 99 629C doi 10 1016 0022 5193 82 90191 6 ISSN 0022 5193 PMID 7183857 Schauperl M Podewitz M Waldner BJ Liedl KR 2016 Enthalpic and Entropic Contributions to Hydrophobicity Journal of Chemical Theory and Computation 12 9 4600 10 doi 10 1021 acs jctc 6b00422 PMC 5024328 PMID 27442443 Eisenberg D July 1984 Three dimensional structure of membrane and surface proteins Annu Rev Biochem 53 595 623 doi 10 1146 annurev bi 53 070184 003115 PMID 6383201 Rose G D Wolfenden R 1993 Hydrogen Bonding Hydrophobicity Packing and Protein Folding Annual Review of Biophysics and Biomolecular Structure Annual Reviews 22 1 381 415 doi 10 1146 annurev bb 22 060193 002121 ISSN 1056 8700 PMID 8347995 a b Bandyopadhyay D Mehler E L 2008 Quantitative expression of protein heterogeneity Response of amino acid side chains to their local environment Proteins Structure Function and Bioinformatics 72 2 646 659 doi 10 1002 prot 21958 PMID 18247345 S2CID 20929779 Hydrophobicity Scales Janin Joel 1979 Surface and inside volumes in globular proteins Nature Springer Science and Business Media LLC 277 5696 491 492 Bibcode 1979Natur 277 491J doi 10 1038 277491a0 ISSN 0028 0836 PMID 763335 S2CID 4338901 Rose G Geselowitz A Lesser G Lee R Zehfus M 1985 08 30 Hydrophobicity of amino acid residues in globular proteins Science American Association for the Advancement of Science AAAS 229 4716 834 838 Bibcode 1985Sci 229 834R doi 10 1126 science 4023714 ISSN 0036 8075 PMID 4023714 a b Wolfenden R Andersson L Cullis P M Southgate C C B 1981 Affinities of amino acid side chains for solvent water Biochemistry American Chemical Society ACS 20 4 849 855 doi 10 1021 bi00507a030 ISSN 0006 2960 PMID 7213619 Y Nozaki C Tanford J Biol Chem 246 1971 2211 Fendler Janos H Nome Faruk Nagyvary Joseph 1975 Compartmentalization of amino acids in surfactant aggregates Journal of Molecular Evolution Springer Science and Business Media LLC 6 3 215 232 Bibcode 1975JMolE 6 215F doi 10 1007 bf01732358 ISSN 0022 2844 PMID 1206727 S2CID 2394979 Leodidis Epaminondas B Hatton T Alan 1990 Amino acids in AOT reversed micelles 2 The hydrophobic effect and hydrogen bonding as driving forces for interfacial solubilization The Journal of Physical Chemistry American Chemical Society ACS 94 16 6411 6420 doi 10 1021 j100379a047 ISSN 0022 3654 Sharp Kim A Nicholls Anthony Friedman Richard Honig Barry 1991 10 08 Extracting hydrophobic free energies from experimental data relationship to protein folding and theoretical models Biochemistry American Chemical Society ACS 30 40 9686 9697 doi 10 1021 bi00104a017 ISSN 0006 2960 PMID 1911756 Zaslavsky B Yu Mestechkina N M Miheeva L M Rogozhin S V 1982 Measurement of relative hydrophobicity of amino acid side chains by partition in an aqueous two phase polymeric system Hydrophobicity scale for non polar and ionogenic side chains Journal of Chromatography A Elsevier BV 240 1 21 28 doi 10 1016 s0021 9673 01 84003 6 ISSN 0021 9673 S Damadoran K B Song J Biol Chem 261 1986 7220 Ben Naim A 1990 02 15 Solvent effects on protein association and protein folding Biopolymers Wiley 29 3 567 596 doi 10 1002 bip 360290312 ISSN 0006 3525 PMID 2331515 S2CID 25691137 a b Chothia Cyrus 1976 The nature of the accessible and buried surfaces in proteins Journal of Molecular Biology Elsevier BV 105 1 1 12 doi 10 1016 0022 2836 76 90191 1 ISSN 0022 2836 PMID 994183 Moret M A Zebende G F 2007 01 19 Amino acid hydrophobicity and accessible surface area Physical Review E American Physical Society APS 75 1 011920 Bibcode 2007PhRvE 75a1920M doi 10 1103 physreve 75 011920 ISSN 1539 3755 PMID 17358197 Phillips J C 2009 11 20 Scaling and self organized criticality in proteins Lysozymec Physical Review E American Physical Society APS 80 5 051916 Bibcode 2009PhRvE 80e1916P doi 10 1103 physreve 80 051916 ISSN 1539 3755 PMID 20365015 Hodges Robert S Zhu Bing Yan Zhou Nian E Mant Colin T 1994 Reversed phase liquid chromatography as a useful probe of hydrophobic interactions involved in protein folding and protein stability Journal of Chromatography A Elsevier BV 676 1 3 15 doi 10 1016 0021 9673 94 80452 4 ISSN 0021 9673 PMID 7921179 Aboderin Akintola A 1971 An empirical hydrophobicity scale for a amino acids and some of its applications International Journal of Biochemistry Elsevier BV 2 11 537 544 doi 10 1016 0020 711x 71 90023 1 ISSN 0020 711X Pliska Vladimir Schmidt Manfred Fauchere Jean Luc 1981 Partition coefficients of amino acids and hydrophobic parameters p of their side chains as measured by thin layer chromatography Journal of Chromatography A Elsevier BV 216 79 92 doi 10 1016 s0021 9673 00 82337 7 ISSN 0021 9673 Plass Monika Valko Klara Abraham Michael H 1998 Determination of solute descriptors of tripeptide derivatives based on high throughput gradient high performance liquid chromatography retention data Journal of Chromatography A Elsevier BV 803 1 2 51 60 doi 10 1016 s0021 9673 97 01215 6 ISSN 0021 9673 Yutani K Ogasahara K Tsujita T Sugino Y 1987 07 01 Dependence of conformational stability on hydrophobicity of the amino acid residue in a series of variant proteins substituted at a unique position of tryptophan synthase alpha subunit Proceedings of the National Academy of Sciences USA Proceedings of the National Academy of Sciences 84 13 4441 4444 Bibcode 1987PNAS 84 4441Y doi 10 1073 pnas 84 13 4441 ISSN 0027 8424 PMC 305105 PMID 3299367 Bull Henry B Breese Keith 1974 Surface tension of amino acid solutions A hydrophobicity scale of the amino acid residues Archives of Biochemistry and Biophysics Elsevier BV 161 2 665 670 doi 10 1016 0003 9861 74 90352 x ISSN 0003 9861 PMID 4839053 Eisenberg David McLachlan Andrew D 1986 Solvation energy in protein folding and binding Nature Springer Science and Business Media LLC 319 6050 199 203 Bibcode 1986Natur 319 199E doi 10 1038 319199a0 ISSN 0028 0836 PMID 3945310 S2CID 21867582 Palliser Christopher C Parry David A D 2000 Quantitative comparison of the ability of hydropathy scales to recognize surface strands in proteins Proteins Structure Function and Genetics Wiley 42 2 243 255 doi 10 1002 1097 0134 20010201 42 2 lt 243 aid prot120 gt 3 0 co 2 b ISSN 0887 3585 PMID 11119649 S2CID 23839522 G Trinquier Y H Sanejouand Protein Eng 11 1998 153 a b c White Stephen 2006 06 29 Experimentally Determined Hydrophobicity Scales University of California Irvine Retrieved 2009 06 12 Wimley William C White Stephen H 1996 Experimentally determined hydrophobicity scale for proteins at membrane interfaces Nature Structural amp Molecular Biology Springer Science and Business Media LLC 3 10 842 848 doi 10 1038 nsb1096 842 ISSN 1545 9993 PMID 8836100 S2CID 1823375 Wimley William C Creamer Trevor P White Stephen H 1996 Solvation Energies of Amino Acid Side Chains and Backbone in a Family of Host Guest Pentapeptides Biochemistry American Chemical Society ACS 35 16 5109 5124 doi 10 1021 bi9600153 ISSN 0006 2960 PMID 8611495 White SH amp Wimley WC 1998 Biochim Biophys Acta 1376 339 352 White Stephen H Wimley William C 1999 MEMBRANE PROTEIN FOLDING AND STABILITY Physical Principles Annual Review of Biophysics and Biomolecular Structure Annual Reviews 28 1 319 365 doi 10 1146 annurev biophys 28 1 319 ISSN 1056 8700 PMID 10410805 Zhu Chongqin Gao Yurui Li Hui Meng Sheng Li Lei Francisco Joseph S Zeng Xiao Cheng 2016 Characterizing hydrophobicity of amino acid side chains in a protein environment via measuring contact angle of a water nanodroplet on planar peptide network Proceedings of the National Academy of Sciences 113 46 12946 12951 doi 10 1073 pnas 1616138113 PMC 5135335 PMID 27803319 Godawat R Jamadagni S N Garde S 2009 Characterizing hydrophobicity of interfaces by using cavity formation solute binding and water correlations Proceedings of the National Academy of Sciences 106 36 15119 15124 doi 10 1073 pnas 0902778106 PMC 2741215 PMID 19706896 External links EditProtScale web based tool for calculating hydropathy plots NetSurfP Secondary Structure and Surface accessibility predictor Whole residue hydrophobicity scale Membrane protein explorer Retrieved from https en wikipedia org w index php title Hydrophobicity scales amp oldid 1141305828, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.