Cross Link

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Cross-link

In chemistry and biology a cross-link is a bond that links one polymer chain to
another. These links may take the form of covalent bonds or ionic bonds and the
polymers can be either synthetic polymers or natural polymers (such as
proteins).

In polymer chemistry "cross-linking" usually refers to the use of cross-links to


promote a change in the polymers' physical properties.

When "crosslinking" is used in the biological field, it refers to the use of a probe
to link proteins together to check for protein–protein interactions, as well as
other creative cross-linking methodologies.
Vulcanization is an example of cross-
Although the term is used to refer to the "linking of polymer chains" for both linking. Schematic presentation of
sciences, the extent of crosslinking and specificities of the crosslinking agents two "polymer chains" (blue and
vary greatly. As with all science, there are overlaps, and the following green) cross-linked after the
delineations are a starting point to understanding the subtleties. vulcanization of natural rubber with
sulfur (n = 0, 1, 2, 3 …).

Contents IUPAC definition


Polymer chemistry A small region in a macromolecule from which at least four chains
Formation emanate, and formed by reactions involving sites or groups on existing
macromolecules or by interactions between existing macromolecules.
Physical cross-links
Oxidative cross-links Notes

In biology 1. The small region may be an atom, a group of atoms, or a number of


Use in protein study
branch points connected by bonds, groups of atoms, or oligomeric chains.
Uses for crosslinked polymers
Measuring degree of crosslinking 2. In the majority of cases, a crosslink is a covalent structure but the term
is also used to describe sites of weaker chemical interactions, portions of
See also crystallites, and even physical interactions and entanglements.[1]
References
External links

Polymer chemistry
Crosslinking is the general term for the process of forming covalent bonds or relatively short sequences of chemical bonds to join
two polymer chains together. The term curing refers to the crosslinking of thermosetting resins, such as unsaturated polyester and
epoxy resin, and the term vulcanization is characteristically used for rubbers.[2] When polymer chains are crosslinked, the
material becomes more rigid.

In polymer chemistry, when a synthetic polymer is said to be "cross-linked", it usually means that the entire bulk of the polymer
has been exposed to the cross-linking method. The resulting modification of mechanical properties depends strongly on the cross-
link density. Low cross-link densities decrease the viscosities of polymer melts. Intermediate cross-link densities transform
gummy polymers into materials that have elastomeric properties and potentially high strengths. Very high cross-link densities can
cause materials to become very rigid or glassy, such as phenol-formaldehyde materials.[3]

Formation
Cross-links can be formed by chemical reactions that
are initiated by heat, pressure, change in pH, or
Typical vinyl ester resin derived from bisphenol A diglycidyl irradiation. For example, mixing of an unpolymerized
ether. Free-radical polymerization gives a highly crosslinked or partially polymerized resin with specific chemicals
polymer.[4] called crosslinking reagents results in a chemical
reaction that forms cross-links. Cross-linking can also
be induced in materials that are normally thermoplastic
through exposure to a radiation source, such as electron beam exposure,[5] gamma radiation, or UV light. For example, electron
beam processing is used to cross-link the C type of cross-linked polyethylene. Other types of cross-linked polyethylene are made
by addition of peroxide during extruding (type A) or by addition of a cross-linking agent (e.g. vinylsilane) and a catalyst during
extruding and then performing a post-extrusion curing.

The chemical process of vulcanization is a type of cross-linking that changes rubber to the hard, durable material associated with
car and bike tires. This process is often called sulfur curing; the term vulcanization comes from Vulcan, the Roman god of fire.
This is, however, a slower process. A typical car tire is cured for 15 minutes at 150 °C. However, the time can be reduced by the
addition of accelerators such as 2-benzothiazolethiol or tetramethylthiuram disulfide. Both of these contain a sulfur atom in the
molecule that initiates the reaction of the sulfur chains with the rubber. Accelerators increase the rate of cure by catalysing the
addition of sulfur chains to the rubber molecules.

Cross-links are the characteristic property of thermosetting plastic materials. In most cases, cross-linking is irreversible, and the
resulting thermosetting material will degrade or burn if heated, without melting. Especially in the case of commercially used
plastics, once a substance is cross-linked, the product is very hard or impossible to recycle. In some cases, though, if the cross-
link bonds are sufficiently different, chemically, from the bonds forming the polymers, the process can be reversed. Permanent
wave solutions, for example, break and re-form naturally occurring cross-links (disulfide bonds) between protein chains in hair.

Physical cross-links
Where chemical cross-links are covalent bonds, physical cross-links are formed by weak interactions. For example, sodium
alginate gels upon exposure to calcium ion, which allows it to form ionic bonds that bridge between alginate chains.[6] Polyvinyl
alcohol gels upon the addition of borax through hydrogen bonding between boric acid and the polymer's alcohol groups.[7] [8]

Other examples of materials which form physically cross-linked gels include gelatin, collagen, agarose, and agar agar.

Chemical covalent cross-links are stable mechanically and thermally, so once formed are difficult to break. Therefore, cross-
linked products like car tires cannot be recycled easily. A class of polymers known as thermoplastic elastomers rely on physical
cross-links in their microstructure to achieve stability, and are widely used in non-tire applications, such as snowmobile tracks,
and catheters for medical use. They offer a much wider range of properties than conventional cross-linked elastomers because the
domains that act as cross-links are reversible, so can be reformed by heat. The stabilizing domains may be non-crystalline (as in
styrene-butadiene block copolymers) or crystalline as in thermoplastic copolyesters.

Note: A rubber which cannot be reformed by heat or chemical treatment is called a thermoset elastomer. On the other hand, a
thermoplastic elastomer can be molded and recycled by heat.

Oxidative cross-links
Many polymers undergo oxidative cross-linking, typically when exposed to
atmospheric oxygen. In some cases this is undesirable and thus polymerization
reactions may involve the use of an antioxidant to slow the formation of
oxidative cross-links. In other cases, when formation of cross-links by oxidation
is desirable, an oxidizer such as hydrogen peroxide may be used to speed up the
process. The compound
bis(triethoxysilylpropyl)tetrasulfide is
The aforementioned process of applying a permanent wave to hair is one a cross-linking agent: the siloxy
example of oxidative cross-linking. In that process the disulfide bonds are groups link to silica and the
reduced, typically using a mercaptan such as ammonium thioglycolate. polysulfide groups vulcanize with
polyolefins.
Following this, the hair is curled and then "neutralized". The neutralizer is
typically an acidic solution of hydrogen peroxide, which causes new disulfide
bonds to form under conditions of oxidation, thus permanently fixing the hair into its new configuration.

In biology
Proteins naturally present in the body can contain crosslinks generated by enzyme-catalyzed or spontaneous reactions. Such
crosslinks are important in generating mechanically stable structures such as hair, skin, and cartilage. Disulfide bond formation is
one of the most common crosslinks, but isopeptide bond formation is also common. Proteins can also be cross-linked artificially
using small-molecule crosslinkers. Compromised collagen in the cornea, a condition known as keratoconus, can be treated with
clinical crosslinking.[9]

In biological context crosslinking could play a role in atheroscelerosis through AGEs which have been implicated to induce
crosslinking of collagen which may lead to vascular stiffening.[10]

Use in protein study


The interactions or mere proximity of proteins can be studied by the clever use of crosslinking agents. For example, protein A and
protein B may be very close to each other in a cell, and a chemical crosslinker[11] could be used to probe the protein–protein
interaction between these two proteins by linking them together, disrupting the cell, and looking for the crosslinked proteins.[12]

A variety of crosslinkers are used to analyze subunit structure of proteins, protein interactions, and various parameters of protein
function by using differing crosslinkers, often with diverse spacer arm lengths. Subunit structure is deduced, since crosslinkers
bind only surface residues in relatively close proximity in the native state. Protein interactions are often too weak or transient to
be easily detected, but by crosslinking, the interactions can be stabilized, captured, and analyzed.

Examples of some common crosslinkers are the imidoester crosslinker dimethyl suberimidate, the N-Hydroxysuccinimide-ester
crosslinker BS3 and formaldehyde. Each of these crosslinkers induces nucleophilic attack of the amino group of lysine and
subsequent covalent bonding via the crosslinker. The zero-length carbodiimide crosslinker EDC functions by converting
carboxyls into amine-reactive isourea intermediates that bind to lysine residues or other available primary amines. SMCC or its
water-soluble analog, Sulfo-SMCC, is commonly used to prepare antibody-hapten conjugates for antibody development.

In-vitro cross-linking method, termed PICUP (photo-induced cross-linking of unmodified proteins), was developed in 1999.[13]
They devised a process in which ammonium persulfate (APS), which acts as an electron acceptor, and tris-bipyridylruthenium (II)
cation ([Ru(bpy)3]2+) are added to the protein of interest and irradiated with UV light.[13] PICUP is more expeditious and high
yielding compared to previous chemical cross-linking methods.[13]

In-vivo crosslinking of protein complexes using photo-reactive amino acid analogs was introduced in 2005 by researchers from
the Max Planck Institute of Molecular Cell Biology and Genetics.[14] In this method, cells are grown with photoreactive diazirine
analogs to leucine and methionine, which are incorporated into proteins. Upon exposure to ultraviolet light, the diazirines are
activated and bind to interacting proteins that are within a few ångströms of the photo-reactive amino acid analog (UV cross-
linking).

Uses for crosslinked polymers


Synthetically crosslinked polymers have many uses, including those in the biological sciences, such as applications in forming
polyacrylamide gels for gel electrophoresis. Synthetic rubber used for tires is made by crosslinking rubber through the process of
vulcanization. This crosslinking makes them more elastic. Hard-shell kayaks are also often manufactured with crosslinked
polymers.

Other examples of polymers that can be crosslinked are ethylene-vinyl acetate–as used in solar panel manufacturing[15] – and
polyethylene.[16][17][18]

Alkyd enamels, the dominant type of commercial oil-based paint, cure by oxidative crosslinking after exposure to air.

In many hydraulic fracturing treatments, a delayed gel-cross-linker fluid is used to carry out fracture treatment of the rock.

The earliest examples of crosslinking, linking long chains of polymers together to increase strength and mass, involved tires.
Rubber was vulcanized with sulfur under heat, which created a link between latex models.[19]

Novel uses for crosslinking can be found in regenerative medicine, where bio-scaffolds are crosslinked to improve their
mechanical properties.[20] More specifically increasing the resistance to dissolution in water based solutions.

Measuring degree of crosslinking


Crosslinking is often measured by swelling tests. The crosslinked sample is placed into a good solvent at a specific temperature,
and either the change in mass or the change in volume is measured. The more crosslinking, the less swelling is attainable. Based
on the degree of swelling, the Flory Interaction Parameter (which relates the solvent interaction with the sample), and the density
of the solvent, the theoretical degree of crosslinking can be calculated according to Flory's Network Theory.[21] Two ASTM
standards are commonly used to describe the degree of crosslinking in thermoplastics. In ASTM D2765, the sample is weighed,
then placed in a solvent for 24 hours, weighed again while swollen, then dried and weighed a final time.[22] The degree of
swelling and the soluble portion can be calculated. In another ASTM standard, F2214, the sample is placed in an instrument that
measures the height change in the sample, allowing the user to measure the volume change.[23] The crosslink density can then be
calculated.

See also
Branching (polymer chemistry)
Cross-linked enzyme aggregate
Cross-linked polyethylene (PEX)
Crosslinking of DNA
Fixation (histology)
Phenol formaldehyde resin (phenolic resin)

References
1. Jenkins, A. D. (1996). "Glossary of basic terms in polymer science (IUPAC Recommendations 1996)" (http://ww
w.iupac.org/publications/pac/1996/pdf/6812x2287.pdf) (PDF). Pure and Applied Chemistry. 68 (12): 2287–2311.
doi:10.1351/pac199668122287 (https://doi.org/10.1351%2Fpac199668122287).
2. Hans Zweifel; Ralph D. Maier; Michael Schiller (2009). Plastics additives handbook (6th ed.). Munich: Hanser.
p. 746. ISBN 978-3-446-40801-2.
3. Gent, Alan N. (1 April 2018). Engineering with Rubber: How to Design Rubber Components (https://books.googl
e.com/books?id=q034u2kLAagC&pg=PA22). Hanser. ISBN 9781569902998. Retrieved 1 April 2018 – via Google
Books.
4. Pham, Ha Q.; Marks, Maurice J. (2012). Epoxy Resins. Ullmann's Encyclopedia of Industrial Chemistry.
doi:10.1002/14356007.a09_547.pub2 (https://doi.org/10.1002%2F14356007.a09_547.pub2). ISBN 978-
3527306732.
5. "Shrink Wrap" (https://www.symmetrymagazine.org/article/october-2009/accelerator-application-shrink-wrap).
symmetry magazine. Retrieved 28 December 2017.
6. Hecht, Hadas; Srebnik, Simcha (2016). "Structural Characterization of Sodium Alginate and Calcium Alginate".
Biomacromolecules. 17 (6): 2160–2167. doi:10.1021/acs.biomac.6b00378 (https://doi.org/10.1021%2Facs.bioma
c.6b00378). PMID 27177209 (https://www.ncbi.nlm.nih.gov/pubmed/27177209).
7. http://www.rsc.org/learn-chemistry/content/filerepository/CMP/00/000/835/cfns%20experiment%2076%20-
%20pva%20polymer%20slime.pdf
8. Casassa, E.Z; Sarquis, A.M; Van Dyke, C.H (1986). "The gelation of polyvinyl alcohol with borax: A novel class
participation experiment involving the preparation and properties of a "slime" ". Journal of Chemical Education. 63
(1): 57. Bibcode:1986JChEd..63...57C (https://ui.adsabs.harvard.edu/abs/1986JChEd..63...57C).
doi:10.1021/ed063p57 (https://doi.org/10.1021%2Fed063p57).
9. Wollensak G, Spoerl E, Seiler T. Riboflavin/ultraviolet-a-induced collagen crosslinking for the treatment of
keratoconus. Am J Ophthalmol. 2003 May;135(5):620-7.
10. Prasad, Anand; Bekker, Peter; Tsimikas, Sotirios (2012-08-01). "Advanced glycation end products and diabetic
cardiovascular disease". Cardiology in Review. 20 (4): 177–183. doi:10.1097/CRD.0b013e318244e57c (https://d
oi.org/10.1097%2FCRD.0b013e318244e57c). ISSN 1538-4683 (https://www.worldcat.org/issn/1538-4683).
PMID 22314141 (https://www.ncbi.nlm.nih.gov/pubmed/22314141).
11. "Pierce Protein Biology - Thermo Fisher Scientific" (http://www.piercenet.com/Objects/View.cfm?type=Page&ID=
FE7F690D-58AE-4342-AE85-BA94DCA642F8). www.piercenet.com. Retrieved 1 April 2018.
12. Kou Qin; Chunmin Dong; Guangyu Wu; Nevin A Lambert (August 2011). "Inactive-state preassembly of Gq-
coupled receptors and Gq heterotrimers" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3177959). Nature
Chemical Biology. 7 (11): 740–747. doi:10.1038/nchembio.642 (https://doi.org/10.1038%2Fnchembio.642).
PMC 3177959 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3177959). PMID 21873996 (https://www.ncbi.nlm.
nih.gov/pubmed/21873996).
13. Fancy, David A.; Kodadek, Thomas (1999-05-25). "Chemistry for the analysis of protein–protein interactions:
Rapid and efficient cross-linking triggered by long wavelength light" (https://www.ncbi.nlm.nih.gov/pmc/articles/P
MC26828). Proceedings of the National Academy of Sciences. 96 (11): 6020–6024.
Bibcode:1999PNAS...96.6020F (https://ui.adsabs.harvard.edu/abs/1999PNAS...96.6020F).
doi:10.1073/pnas.96.11.6020 (https://doi.org/10.1073%2Fpnas.96.11.6020). ISSN 0027-8424 (https://www.world
cat.org/issn/0027-8424). PMC 26828 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC26828). PMID 10339534 (ht
tps://www.ncbi.nlm.nih.gov/pubmed/10339534).
14. Suchanek, Monika; Anna Radzikowska; Christoph Thiele (April 2005). "Photo-leucine and photo-methionine allow
identification of protein–protein interactions in living cells". Nature Methods. 2 (4): 261–268.
doi:10.1038/nmeth752 (https://doi.org/10.1038%2Fnmeth752). PMID 15782218 (https://www.ncbi.nlm.nih.gov/pu
bmed/15782218).
15. solar cell manufacturing and solar panel production (https://www.youtube.com/watch?time_continue=341&v=fZ1
SC-vUe_I), at 3:25, is "a layer of EVA", and another.... later bonded around the solar cells themselves under heat
and pressure, "causing cross-linking of the EVA to form a chemical bond which hermetically seals the module.",
accessed 4 September 2018.
16. Reyes-Labarta, J.A.; Marcilla, A. (2012). "Thermal Treatment and Degradation of Crosslinked Ethylene Vinyl
Acetate-Polyethylene-Azodicarbonamide-ZnO Foams. Complete Kinetic Modelling and Analysis". Industrial &
Engineering Chemistry Research. 51 (28): 9515–9530. doi:10.1021/ie3006935 (https://doi.org/10.1021%2Fie300
6935).
17. Reyes-Labarta, J.A.; Marcilla, A.; Sempere, J. (2011). "Kinetic Study of the Thermal Processing and Pyrolysis of
Crosslinked Ethylene Vinyl Acetate-Polyethylene Mixtures". Industrial & Engineering Chemistry Research. 50
(13): 7964–7976. doi:10.1021/ie200276v (https://doi.org/10.1021%2Fie200276v).
18. Reyes-Labarta, J.A.; Olaya, M.M.; Marcilla, A. (2006). "DSC and TGA Study of the Transitions Involved in the
Thermal Treatment of Binary Mixtures of PE and EVA Copolymer with a Crosslinking Agent". Polymer. 47 (24):
8194–8202. doi:10.1016/j.polymer.2006.09.054 (https://doi.org/10.1016%2Fj.polymer.2006.09.054).
19. "Crosslinking and Tires" (http://www.ebeam.com/blog/tire-perfection-with-electron-beam). ebeam.com. Retrieved
1 April 2018.
20. Lien, S.-M.; Li, W.-T.; Huang, T.-J. (2008). "Genipin-crosslinked gelatin scaffolds for articular cartilage tissue
engineering with a novel crosslinking method". Materials Science and Engineering: C. 28 (1): 36–43.
doi:10.1016/j.msec.2006.12.015 (https://doi.org/10.1016%2Fj.msec.2006.12.015).
21. Flory, P.J., "Principles of Polymer Chemistry" (1953)
22. "ASTM D2765 - 16 Standard Test Methods for Determination of Gel Content and Swell Ratio of Crosslinked
Ethylene Plastics" (http://www.astm.org/Standards/D2765.htm). www.astm.org. Retrieved 1 April 2018.
23. "ASTM F2214 - 16 Standard Test Method for In Situ Determination of Network Parameters of Crosslinked Ultra
High Molecular Weight Polyethylene (UHMWPE)" (http://www.astm.org/Standards/F2214.htm). www.astm.org.
Retrieved 1 April 2018.

External links
Application note on how to measure degree of crosslinking in plastics (http://www.campoly.com/index.php/downlo
ad_file/view/380/108/)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Cross-link&oldid=913906862"

This page was last edited on 3 September 2019, at 23:34 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; 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.

You might also like