Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Steric hindrance  





2 Measures of steric properties  



2.1  Rate data  





2.2  A-values  





2.3  Ceiling temperatures  





2.4  Cone angles  







3 Significance and applications  





4 See also  





5 References  





6 External links  














Steric effects






العربية

Català
Čeština
Deutsch
Eesti
Español
فارسی
Français

Bahasa Indonesia
Italiano
עברית
Nederlands

Oʻzbekcha / ўзбекча
Polski
Português
Română
Русский
Simple English
Српски / srpski
Srpskohrvatski / српскохрватски
Suomi
Svenska
Українська

 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 




In other projects  



Wikimedia Commons
 
















Appearance
   

 






From Wikipedia, the free encyclopedia
 

(Redirected from Sterically)

The parent cyclobutadiene (R = H) readily dimerizes but the R = tert-butyl derivative is robust.[1]

Steric effects arise from the spatial arrangement of atoms. When atoms come close together there is generally a rise in the energy of the molecule. Steric effects are nonbonding interactions that influence the shape (conformation) and reactivity of ions and molecules. Steric effects complement electronic effects, which dictate the shape and reactivity of molecules. Steric repulsive forces between overlapping electron clouds result in structured groupings of molecules stabilized by the way that opposites attract and like charges repel.

Steric hindrance[edit]

Regioselective dimethoxytritylation of the primary 5'-hydroxyl group of thymidine in the presence of a free secondary 3'-hydroxy group as a result of steric hindrance due to the dimethoxytrityl group and the ribose ring (Py = pyridine).[2]

Steric hindrance is a consequence of steric effects. Steric hindrance is the slowing of chemical reactions due to steric bulk. It is usually manifested in intermolecular reactions, whereas discussion of steric effects often focus on intramolecular interactions. Steric hindrance is often exploited to control selectivity, such as slowing unwanted side-reactions.

Steric hindrance between adjacent groups can also affect torsional bond angles. Steric hindrance is responsible for the observed shape of rotaxanes and the low rates of racemization of 2,2'-disubstituted biphenyl and binaphthyl derivatives.

Measures of steric properties[edit]

Because steric effects have profound impact on properties, the steric properties of substituents have been assessed by numerous methods.

Rate data[edit]

Relative rates of chemical reactions provide useful insights into the effects of the steric bulk of substituents. Under standard conditions, methyl bromide solvolyzes107 faster than does neopentyl bromide. The difference reflects the inhibition of attack on the compound with the sterically bulky (CH3)3C group.[3]

A-values[edit]

A-values provide another measure of the bulk of substituents. A-values are derived from equilibrium measurements of monosubstituted cyclohexanes.[4][5][6][7] The extent that a substituent favors the equatorial position gives a measure of its bulk.

The A-value for a methyl group is 1.74 as derived from the chemical equilibrium above. It costs 1.74 kcal/mol for the methyl group to adopt to the axial position compared to the equatorial position.
Substituent A-Value
H 0
CH3 1.74
CH2CH3 1.75
CH(CH3)2 2.15
C(CH3)3 >4

Ceiling temperatures[edit]

Ceiling temperature () is a measure of the steric properties of the monomers that comprise a polymer. is the temperature where the rate of polymerization and depolymerization are equal. Sterically hindered monomers give polymers with low 's, which are usually not useful.

Monomer Ceiling temperature (°C)[8] Structure
ethylene 610 CH2=CH2
isobutylene 175 CH2=CMe2
1,3-butadiene 585 CH2=CHCH=CH2
isoprene 466 CH2=C(Me)CH=CH2
styrene 395 PhCH=CH2
α-methylstyrene 66 PhC(Me)=CH2

Cone angles[edit]

Ligand cone angle.

Ligand cone angles are measures of the size of ligandsincoordination chemistry. It is defined as the solid angle formed with the metal at the vertex and the hydrogen atoms at the perimeter of the cone (see figure).[9]

Cone angles of common phosphine ligands
Ligand Angle (°)
PH3 87
P(OCH3)3 107
P(CH3)3 118
P(CH2CH3)3 132
P(C6H5)3 145
P(cyclo-C6H11)3 179
P(t-Bu)3 182
P(2,4,6-Me3C6H2)3 212

Significance and applications[edit]

Steric effects are critical to chemistry, biochemistry, and pharmacology. In organic chemistry, steric effects are nearly universal and affect the rates and activation energies of most chemical reactions to varying degrees.

In biochemistry, steric effects are often exploited in naturally occurring molecules such as enzymes, where the catalytic site may be buried within a large protein structure. In pharmacology, steric effects determine how and at what rate a drug will interact with its target bio-molecules.

The steric effect of tri-(tert-butyl)amine makes electrophilic reactions, like forming the tetraalkylammonium cation, difficult. It is difficult for electrophiles to get close enough to allow attack by the lone pair of the nitrogen (nitrogen is shown in blue)

See also[edit]

References[edit]

  1. ^ Günther Maier; Stephan Pfriem; Ulrich Schäfer; Rudolf Matusch (1978). "Tetra-tert-butyltetrahedrane". Angew. Chem. Int. Ed. Engl. 17 (7): 520–1. doi:10.1002/anie.197805201.
  • ^ Gait, Michael (1984). Oligonucleotide synthesis: a practical approach. Oxford: IRL Press. ISBN 0-904147-74-6.
  • ^ Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, ISBN 978-0-471-72091-1
  • ^ E.L. Eliel, S.H. Wilen and L.N. Mander, Stereochemistry of Organic Compounds, Wiley, New York (1994). ISBN 81-224-0570-3
  • ^ Eliel, E.L.; Allinger, N.L.; Angyal, S.J.; G.A., Morrison (1965). Conformational Analysis. New York: Interscience Publishers.
  • ^ Hirsch, J.A. (1967). Topics in Stereochemistry (first ed.). New York: John Wiley & Sons, Inc. p. 199.
  • ^ Romers, C.; Altona, C.; Buys, H.R.; Havinga, E. (1969). Topics in Stereochemistry (fourth ed.). New York: John Wiley & Sons, Inc. p. 40.
  • ^ Stevens, Malcolm P. (1999). "6". Polymer Chemistry an Introduction (3rd ed.). New York: Oxford University Press. pp. 193–194. ISBN 978-0-19-512444-6.
  • ^ Tolman, Chadwick A. (1970-05-01). "Phosphorus ligand exchange equilibriums on zerovalent nickel. Dominant role for steric effects". J. Am. Chem. Soc. 92 (10): 2956–2965. doi:10.1021/ja00713a007.
  • ^ Stephan, Douglas W. "Frustrated Lewis pairs": a concept for new reactivity and catalysis. Org. Biomol. Chem. 2008, 6, 1535–1539. doi:10.1039/b802575b
  • ^ Helmut Fiege; Heinz-Werner Voges; Toshikazu Hamamoto; Sumio Umemura; Tadao Iwata; Hisaya Miki; Yasuhiro Fujita; Hans-Josef Buysch; Dorothea Garbe; Wilfried Paulus (2002). "Phenol Derivatives". Ullmann's Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. pp. a19_313. doi:10.1002/14356007.a19_313. ISBN 3-527-30673-0.
  • ^ Pieter Gijsman (2010). "Photostabilisation of Polymer Materials". In Norman S. Allen (ed.). Photochemistry and Photophysics of Polymer Materials Photochemistry. Hoboken: John Wiley & Sons. pp. 627–679. doi:10.1002/9780470594179.ch17. ISBN 978-0-470-59417-9..
  • ^ Klaus Köhler; Peter Simmendinger; Wolfgang Roelle; Wilfried Scholz; Andreas Valet; Mario Slongo (2010). "Paints and Coatings, 4. Pigments, Extenders, and Additives". Ullmann's Encyclopedia Of Industrial Chemistry. pp. o18_o03. doi:10.1002/14356007.o18_o03. ISBN 978-3-527-30673-2.
  • ^ Goto, Kei; Nagahama, Michiko; Mizushima, Tadashi; Shimada, Keiichi; Kawashima, Takayuki; Okazaki, Renji (2001). "The First Direct Oxidative Conversion of a Selenol to a Stable Selenenic Acid: Experimental Demonstration of Three Processes Included in the Catalytic Cycle of Glutathione Peroxidase". Organic Letters. 3 (22): 3569–3572. doi:10.1021/ol016682s. PMID 11678710.
  • External links[edit]


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Steric_effects&oldid=1224855863"

    Categories: 
    Stereochemistry
    Physical organic chemistry
    Hidden categories: 
    Articles with short description
    Short description matches Wikidata
    Webarchive template wayback links
     



    This page was last edited on 20 May 2024, at 21:46 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki