To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Natural rubber

From Wikipedia, the free encyclopedia

Photo of pieces of natural rubber in a glass jar.
Pieces of natural vulcanized rubber at Hutchinson's Research and Innovation Center in France.
Latex being collected from a tapped rubber tree, Cameroon
Rubber tree plantation in Thailand

Rubber, also called India rubber, latex, Amazonian rubber, caucho, or caoutchouc,[1] as initially produced, consists of polymers of the organic compound isoprene, with minor impurities of other organic compounds. Thailand, Malaysia, Indonesia, and Cambodia are four of the leading rubber producers.[2][3][4]

Types of polyisoprene that are used as natural rubbers are classified as elastomers.

Currently, rubber is harvested mainly in the form of the latex from the Pará rubber tree (Hevea brasiliensis) or others. The latex is a sticky, milky and white colloid drawn off by making incisions in the bark and collecting the fluid in vessels in a process called "tapping". The latex then is refined into the rubber that is ready for commercial processing. In major areas, latex is allowed to coagulate in the collection cup. The coagulated lumps are collected and processed into dry forms for sale.

Natural rubber is used extensively in many applications and products, either alone or in combination with other materials. In most of its useful forms, it has a large stretch ratio and high resilience and also is water-proof.[citation needed]

Industrial demand for rubber-like materials began to outstrip natural rubber supplies by the end of the 19th century, leading to the synthesis of synthetic rubber in 1909 by chemical means.[citation needed]

YouTube Encyclopedic

  • 1/5
    Views:
    237 008
    3 238
    628 588
    17 027
    2 303
  • Incredible production process from natural rubber to giant car tires
  • 🔥what is Vulcanization? Natural Rubber | Polymer #shorts #reels #organicchemistry #jee #neet
  • How natural Rubber is Made from Trees | Rubber Harvesting and Processing | Rubber Tapping Method
  • Natural Rubber - Polymers - Chemistry Class 12
  • Natural Rubber Latex - Polymers - Engineering Chemistry 1

Transcription

Varieties

Amazonian rubber tree (Hevea brasiliensis)

The major commercial source of natural rubber latex is the Amazonian rubber tree (Hevea brasiliensis),[1] a member of the spurge family, Euphorbiaceae. Once native to Brazil, the species is now pan-tropical. This species is preferred because it grows well under cultivation. A properly managed tree responds to wounding by producing more latex for several years.[citation needed]

Congo rubber (Landolphia owariensis and L. spp.)

Congo rubber, formerly a major source of rubber, which motivated the atrocities in the Congo Free State, came from vines in the genus Landolphia (L. kirkii, L. heudelotis, and L. owariensis).[5]

Dandelion

Dandelion milk contains latex. The latex exhibits the same quality as the natural rubber from rubber trees. In the wild types of dandelion, latex content is low and varies greatly. In Nazi Germany, research projects tried to use dandelions as a base for rubber production, but failed.[6] In 2013, by inhibiting one key enzyme and using modern cultivation methods and optimization techniques, scientists in the Fraunhofer Institute for Molecular Biology and Applied Ecology (IME) in Germany developed a cultivar of the Kazakh dandelion (Taraxacum kok-saghyz) that is suitable for commercial production of natural rubber.[7] In collaboration with Continental Tires, IME began a pilot facility.

Other

Many other plants produce forms of latex rich in isoprene polymers, though not all produce usable forms of polymer as easily as the Pará.[8] Some of them require more elaborate processing to produce anything like usable rubber, and most are more difficult to tap. Some produce other desirable materials, for example gutta-percha (Palaquium gutta)[9] and chicle from Manilkara species. Others that have been commercially exploited, or at least showed promise as rubber sources, include the rubber fig (Ficus elastica), Panama rubber tree (Castilla elastica), various spurges (Euphorbia spp.), lettuce (Lactuca species), the related Scorzonera tau-saghyz, various Taraxacum species, including common dandelion (Taraxacum officinale) and Kazakh dandelion, and, perhaps most importantly for its hypoallergenic properties, guayule (Parthenium argentatum). The term gum rubber is sometimes applied to the tree-obtained version of natural rubber in order to distinguish it from the synthetic version.[10]

History

The first use of rubber was by the indigenous cultures of Mesoamerica. The earliest archeological evidence of the use of natural latex from the Hevea tree comes from the Olmec culture, in which rubber was first used for making balls for the Mesoamerican ballgame. Rubber was later used by the Maya and Aztec cultures – in addition to making balls, Aztecs used rubber for other purposes, such as making containers and to make textiles waterproof by impregnating them with the latex sap.[11][12]

Charles Marie de La Condamine is credited with introducing samples of rubber to the Académie Royale des Sciences of France in 1736.[13] In 1751, he presented a paper by François Fresneau to the Académie (published in 1755) that described many of rubber's properties. This has been referred to as the first scientific paper on rubber.[13] In England, Joseph Priestley, in 1770, observed that a piece of the material was extremely good for rubbing off pencil marks on paper, hence the name "rubber". It slowly made its way around England. In 1764, François Fresnau discovered that turpentine was a rubber solvent. Giovanni Fabbroni is credited with the discovery of naphtha as a rubber solvent in 1779.[citation needed] Charles Goodyear redeveloped vulcanization in 1839, although Mesoamericans had used stabilized rubber for balls and other objects as early as 1600 BC.[14][15]

South America remained the main source of latex rubber used during much of the 19th century. The rubber trade was heavily controlled by business interests but no laws expressly prohibited the export of seeds or plants. In 1876, Henry Wickham smuggled 70,000 Amazonian rubber tree seeds from Brazil and delivered them to Kew Gardens, England. Only 2,400 of these germinated. Seedlings were then sent to India, British Ceylon (Sri Lanka), Dutch East Indies (Indonesia), Singapore, and British Malaya. Malaya (now Peninsular Malaysia) was later to become the biggest producer of rubber.[16]

In the early 1900s, the Congo Free State in Africa was also a significant source of natural rubber latex, mostly gathered by forced labor.[citation needed] King Leopold II's colonial state brutally enforced production quotas. Tactics to enforce the rubber quotas included removing the hands of victims to prove they had been killed. Soldiers often came back from raids with baskets full of chopped-off hands. Villages that resisted were razed to encourage better compliance locally.[citation needed] (See Atrocities in the Congo Free State for more information on the rubber trade in the Congo Free State in the late 1800s and early 1900s.)

The rubber boom in the Amazon also similarly affected indigenous populations to varying degrees. Correrias, or slave raids were frequent in Colombia, Peru and Bolivia where many were either captured or killed. The most well known case of atrocities generated from rubber extraction in South America came from the Putumayo genocide. Between the 1880s–1913 Julio César Arana and his company that would become the Peruvian Amazon Company controlled the Putumayo river. W.E. Hardenburg, Benjamin Saldaña Rocca and Roger Casement were influential figures in exposing these atrocities. Roger Casement was also prominent in revealing the Congo atrocities to the world. Days before entering Iquitos by boat Casement wrote "'Caoutchouc was first called 'india rubber,' because it came from the Indies, and the earliest European use of it was to rub out or erase. It is now called India rubber because it rubs out or erases the Indians."[17][18]

"Enslaved natives with a load of rubber weighing 75 kilos, they have journeyed 100 kilometers with no food given"

In India, commercial cultivation was introduced by British planters, although the experimental efforts to grow rubber on a commercial scale were initiated as early as 1873 at the Calcutta Botanical Garden. The first commercial Hevea plantations were established at Thattekadu in Kerala in 1902. In later years the plantation expanded to Karnataka, Tamil Nadu and the Andaman and Nicobar Islands of India. Today, India is the world's 3rd largest producer and 4th largest consumer of rubber.[19]

In Singapore and Malaya, commercial production was heavily promoted by Sir Henry Nicholas Ridley, who served as the first Scientific Director of the Singapore Botanic Gardens from 1888 to 1911. He distributed rubber seeds to many planters and developed the first technique for tapping trees for latex without causing serious harm to the tree.[20] Because of his fervent promotion of this crop, he is popularly remembered by the nickname "Mad Ridley".[21]

Pre–World War II

Before World War II significant uses included door and window profiles, hoses, belts, gaskets, matting, flooring, and dampeners (antivibration mounts) for the automotive industry. The use of rubber in car tires (initially solid rather than pneumatic) in particular consumed a significant amount of rubber. Gloves (medical, household, and industrial) and toy balloons were large consumers of rubber, although the type of rubber used is concentrated latex. Significant tonnage of rubber was used as adhesives in many manufacturing industries and products, although the two most noticeable were the paper and the carpet industries. Rubber was commonly used to make rubber bands and pencil erasers.

Rubber produced as a fiber, sometimes called 'elastic', had significant value to the textile industry because of its excellent elongation and recovery properties. For these purposes, manufactured rubber fiber was made as either an extruded round fiber or rectangular fibers cut into strips from extruded film. Because of its low dye acceptance, feel and appearance, the rubber fiber was either covered by yarn of another fiber or directly woven with other yarns into the fabric. Rubber yarns were used in foundation garments. While rubber is still used in textile manufacturing, its low tenacity limits its use in lightweight garments because latex lacks resistance to oxidizing agents and is damaged by aging, sunlight, oil and perspiration. The textile industry turned to neoprene (polymer of chloroprene), a type of synthetic rubber, as well as another more commonly used elastomer fiber, spandex (also known as elastane), because of their superiority to rubber in both strength and durability.

Properties

Rubber latex

Rubber exhibits unique physical and chemical properties. Rubber's stress–strain behavior exhibits the Mullins effect and the Payne effect and is often modeled as hyperelastic. Rubber strain crystallizes. Because there are weakened allylic C-H bonds in each repeat unit, natural rubber is susceptible to vulcanisation as well as being sensitive to ozone cracking. The two main solvents for rubber are turpentine and naphtha (petroleum). Because rubber does not dissolve easily, the material is finely divided by shredding prior to its immersion. An ammonia solution can be used to prevent the coagulation of raw latex. Rubber begins to melt at approximately 180 °C (356 °F).

Elasticity

Rubber latex elasticity

On a microscopic scale, relaxed rubber is a disorganized cluster of erratically changing wrinkled chains. In stretched rubber, the chains are almost linear. The restoring force is due to the preponderance of wrinkled conformations over more linear ones. For the quantitative treatment see ideal chain, for more examples see entropic force.

Cooling below the glass transition temperature permits local conformational changes but a reordering is practically impossible because of the larger energy barrier for the concerted movement of longer chains. "Frozen" rubber's elasticity is low and strain results from small changes of bond lengths and angles: this caused the Challenger disaster, when the American Space Shuttle's flattened o-rings failed to relax to fill a widening gap.[22] The glass transition is fast and reversible: the force resumes on heating.

The parallel chains of stretched rubber are susceptible to crystallization. This takes some time because turns of twisted chains have to move out of the way of the growing crystallites. Crystallization has occurred, for example, when, after days, an inflated toy balloon is found withered at a relatively large remaining volume. Where it is touched, it shrinks because the temperature of the hand is enough to melt the crystals.

Vulcanization of rubber creates di- and polysulfide bonds between chains, which limits the degrees of freedom and results in chains that tighten more quickly for a given strain, thereby increasing the elastic force constant and making the rubber harder and less extensible.

Malodour

Raw rubber storage depots and rubber processing can produce malodour that is serious enough to become a source of complaints and protest to those living in the vicinity.[23] Microbial impurities originate during the processing of block rubber. These impurities break down during storage or thermal degradation and produce volatile organic compounds. Examination of these compounds using gas chromatography/mass spectrometry (GC/MS) and gas chromatography (GC) indicates that they contain sulfur, ammonia, alkenes, ketones, esters, hydrogen sulfide, nitrogen, and low-molecular-weight fatty acids (C2–C5).[24][25] When latex concentrate is produced from rubber, sulfuric acid is used for coagulation. This produces malodourous hydrogen sulfide.[25] The industry can mitigate these bad odours with scrubber systems.[25]

Chemical makeup

(1)trans-1,4-polyisoprene is called gutta-percha. (2)in natural rubber various chains are held together by weak Van Der Waal's interactions and has a coiled structure.so it can be stretched like a spring and exhibits elastic properties
Chemical structure of cis-polyisoprene, the main constituent of natural rubber. Synthetic cis-polyisoprene and natural cis-polyisoprene are derived from distinct precursors, isopentenyl pyrophosphate and isoprene.

Rubber is the polymer cis-1,4-polyisoprene – with a molecular weight of 100,000 to 1,000,000 daltons. Typically, a small percentage (up to 5% of dry mass) of other materials, such as proteins, fatty acids, resins, and inorganic materials (salts) are found in natural rubber. Polyisoprene can also be created synthetically, producing what is sometimes referred to as "synthetic natural rubber", but the synthetic and natural routes are distinct.[10] Some natural rubber sources, such as gutta-percha, are composed of trans-1,4-polyisoprene, a structural isomer that has similar properties. Natural rubber is an elastomer and a thermoplastic. Once the rubber is vulcanized, it is a thermoset. Most rubber in everyday use is vulcanized to a point where it shares properties of both; i.e., if it is heated and cooled, it is degraded but not destroyed. The final properties of a rubber item depend not just on the polymer, but also on modifiers and fillers, such as carbon black, factice, whiting and others.

Biosynthesis

Rubber particles are formed in the cytoplasm of specialized latex-producing cells called laticifers within rubber plants.[26] Rubber particles are surrounded by a single phospholipid membrane with hydrophobic tails pointed inward. The membrane allows biosynthetic proteins to be sequestered at the surface of the growing rubber particle, which allows new monomeric units to be added from outside the biomembrane, but within the lacticifer. The rubber particle is an enzymatically active entity that contains three layers of material, the rubber particle, a biomembrane and free monomeric units. The biomembrane is held tightly to the rubber core by the high negative charge along the double bonds of the rubber polymer backbone.[27] Free monomeric units and conjugated proteins make up the outer layer. The rubber precursor is isopentenyl pyrophosphate (an allylic compound), which elongates by Mg2+-dependent condensation by the action of rubber transferase. The monomer adds to the pyrophosphate end of the growing polymer.[citation needed] The process displaces the terminal high-energy pyrophosphate. The reaction produces a cis polymer. The initiation step is catalyzed by prenyltransferase, which converts three monomers of isopentenyl pyrophosphate into farnesyl pyrophosphate.[28] The farnesyl pyrophosphate can bind to rubber transferase to elongate a new rubber polymer.

The required isopentenyl pyrophosphate is obtained from the mevalonate pathway, which derives from acetyl-CoA in the cytosol. In plants, isoprene pyrophosphate can also be obtained from the 1-deox-D-xyulose-5-phosphate/2-C-methyl-D-erythritol-4-phosphate pathway within plasmids.[29] The relative ratio of the farnesyl pyrophosphate initiator unit and isoprenyl pyrophosphate elongation monomer determines the rate of new particle synthesis versus elongation of existing particles. Though rubber is known to be produced by only one enzyme, extracts of latex host numerous small molecular weight proteins with unknown function. The proteins possibly serve as cofactors, as the synthetic rate decreases with complete removal.[30]

Production

Rubber is generally cultivated in large plantations. The image shows a coconut shell used in collecting latex, in plantations in Kerala, India.
Sheets of natural rubber

More than 28 million tons of rubber were produced in 2017, of which approximately 47% was natural. Since the bulk is synthetic, which is derived from petroleum, the price of natural rubber is determined, to a large extent, by the prevailing global price of crude oil.[31][32] Asia was the main source of natural rubber, accounting for about 90% of output in 2021.[33] The three largest producers, Thailand, Indonesia,[34] and Malaysia, together account for around 72% of all natural rubber production. Natural rubber is not cultivated widely in its native continent of South America because of the South American leaf blight, and other natural predators there.

Cultivation

Rubber latex is extracted from rubber trees. The economic life of rubber trees in plantations is around 32 years, with up to 7 years being an immature phase and about 25 years of productive phase.

The soil requirement is well-drained, weathered soil consisting of laterite, lateritic types, sedimentary types, nonlateritic red or alluvial soils.

The climatic conditions for optimum growth of rubber trees are:

  • Rainfall of around 250 centimetres (98 in) evenly distributed without any marked dry season and with at least 100 rainy days per year
  • Temperature range of about 20 to 34 °C (68 to 93 °F), with a monthly mean of 25 to 28 °C (77 to 82 °F)
  • Atmospheric humidity of around 80%
  • About 2,000 hours sunshine per year at the rate of six hours per day throughout the year
  • Absence of strong winds

Many high-yielding clones have been developed for commercial planting. These clones yield more than 2,000 kilograms per hectare (1,800 lb/acre) of dry rubber per year, under ideal conditions.

Collection

Vintage tobacco card, Tapping a Rubber Tree, India, Products of the World series, Player's Cigarettes, 1909

In places such as Kerala and Sri Lanka, where coconuts are in abundance, the half shell of coconut was used as the latex collection container. Glazed pottery or aluminium or plastic cups became more common in Kerala-India and other countries. The cups are supported by a wire that encircles the tree. This wire incorporates a spring so it can stretch as the tree grows. The latex is led into the cup by a galvanised "spout" knocked into the bark. Rubber tapping normally takes place early in the morning, when the internal pressure of the tree is highest. A good tapper can tap a tree every 20 seconds on a standard half-spiral system, and a common daily "task" size is between 450 and 650 trees. Trees are usually tapped on alternate or third days, although many variations in timing, length and number of cuts are used. "Tappers would make a slash in the bark with a small hatchet. These slanting cuts allowed latex to flow from ducts located on the exterior or the inner layer of bark (cambium) of the tree. Since the cambium controls the growth of the tree, growth stops if it is cut. Thus, rubber tapping demanded accuracy, so that the incisions would not be too many given the size of the tree, or too deep, which could stunt its growth or kill it."[35]

A woman in Sri Lanka harvesting rubber, c. 1920

It is usual to tap a panel at least twice, sometimes three times, during the tree's life. The economic life of the tree depends on how well the tapping is carried out, as the critical factor is bark consumption. A standard in Malaysia for alternate daily tapping is 25 cm (vertical) bark consumption per year. The latex-containing tubes in the bark ascend in a spiral to the right. For this reason, tapping cuts usually ascend to the left to cut more tubes. The trees drip latex for about four hours, stopping as latex coagulates naturally on the tapping cut, thus blocking the latex tubes in the bark. Tappers usually rest and have a meal after finishing their tapping work and then start collecting the liquid "field latex" at about midday.

Field coagula

Mixed field coagula.

The four types of field coagula are "cuplump", "treelace", "smallholders' lump", and "earth scrap". Each has significantly different properties.[36] Some trees continue to drip after the collection leading to a small amount of "cup lump" that is collected at the next tapping. The latex that coagulates on the cut is also collected as "tree lace". Tree lace and cup lump together account for 10%–20% of the dry rubber produced. Latex that drips onto the ground, "earth scrap", is also collected periodically for processing of low-grade product.

Cup lump
Cup lump rubber coagula in a Myanmar road stall.

Cup lump is the coagulated material found in the collection cup when the tapper next visits the tree to tap it again. It arises from latex clinging to the walls of the cup after the latex was last poured into the bucket, and from late-dripping latex exuded before the latex-carrying vessels of the tree become blocked. It is of higher purity and of greater value than the other three types.

'Cup lumps' can also be used to describe a completely different type of coagulate that has collected in smallholder plantations over a period of 1–2 weeks. After tapping all of the trees, the tapper will return to each tree and stir in some type of acid, which allows the newly harvested latex to mix with the previously coagulated material. The rubber/acid mixture is what gives rubber plantations, markets, and factories a strong odor.

Tree lace

Tree lace is the coagulum strip that the tapper peels off the previous cut before making a new cut. It usually has higher copper and manganese contents than cup lump. Both copper and manganese are pro-oxidants and can damage the physical properties of the dry rubber.

Smallholders' lump

Smallholders' lump is produced by smallholders, who collect rubber from trees far from the nearest factory. Many Indonesian smallholders, who farm paddies in remote areas, tap dispersed trees on their way to work in the paddy fields and collect the latex (or the coagulated latex) on their way home. As it is often impossible to preserve the latex sufficiently to get it to a factory that processes latex in time for it to be used to make high quality products, and as the latex would anyway have coagulated by the time it reached the factory, the smallholder will coagulate it by any means available, in any container available. Some smallholders use small containers, buckets etc., but often the latex is coagulated in holes in the ground, which are usually lined with plastic sheeting. Acidic materials and fermented fruit juices are used to coagulate the latex — a form of assisted biological coagulation. Little care is taken to exclude twigs, leaves, and even bark from the lumps that are formed, which may also include tree lace.

Earth scrap

Earth scrap is material that gathers around the base of the tree. It arises from latex overflowing from the cut and running down the bark, from rain flooding a collection cup containing latex, and from spillage from tappers' buckets during collection. It contains soil and other contaminants, and has variable rubber content, depending on the amount of contaminants. Earth scrap is collected by field workers two or three times a year and may be cleaned in a scrap-washer to recover the rubber, or sold to a contractor who cleans it and recovers the rubber. It is of low quality.

Processing

Removing coagulum from coagulating troughs.

Latex coagulates in the cups if kept for long and must be collected before this happens. The collected latex, "field latex", is transferred into coagulation tanks for the preparation of dry rubber or transferred into air-tight containers with sieving for ammoniation. Ammoniation, invented by patent lawyer and vice-president of the United States Rubber Company Ernest Hopkinson around 1920, preserves the latex in a colloidal state for longer periods of time. Latex is generally processed into either latex concentrate for manufacture of dipped goods or coagulated under controlled, clean conditions using formic acid. The coagulated latex can then be processed into the higher-grade, technically specified block rubbers such as SVR 3L or SVR CV or used to produce Ribbed Smoke Sheet grades. Naturally coagulated rubber (cup lump) is used in the manufacture of TSR10 and TSR20 grade rubbers. Processing for these grades is a size reduction and cleaning process to remove contamination and prepare the material for the final stage of drying.[37]

The dried material is then baled and palletized for storage and shipment.

Molecular Structure

Rubber is a natural polymer of isoprene (polyisoprene), and an elastomer (a stretchy polymer). Polymers are simply chains of molecules that can be linked together. Rubber is one of the few naturally occurring polymers and prized for its high stretch ratio, resilience, and water-proof properties. Other examples of natural polymers include tortoise shell, amber, and animal horn.[38] When harvested, latex rubber takes the form of latex, an opaque, white, milky suspension of rubber particles in water. It is then transformed through industrial processes to the common solid form so commonly seen today.

Vulcanized rubber

Torn latex rubber dry suit wrist seal

Natural rubber is reactive and vulnerable to oxidization, but it can be stabilized through a heating process called vulcanization. Vulcanization is a process by which the rubber is heated and sulfur, peroxide, or bisphenol are added to improve resistance and elasticity and to prevent it from oxidizing. Carbon black, which can be derived from a petroleum refinery or other natural incineration processes, is sometimes used as an additive to rubber to improve its strength, especially in vehicle tires.[39][40]

During vulcanization, rubber's polyisoprene molecules (long chains of isoprene) are heated and cross-linked with molecular bonds to sulfur, forming a 3-D matrix. The optimal percentage of sulfur is approximately 10%. In this form, the polyisoprene molecules orientation is still random but they become aligned when the rubber is stretched. This sulfur vulcanization makes the rubber stronger and more rigid, but still very elastic.[41] And through the vulcanization process, the sulfur and latex are meant to be totally used up in individual form.

Transportation

Natural rubber latex is shipped from factories in Southeast Asia, South America, and West and Central Africa to destinations around the world. As the cost of natural rubber has risen significantly and rubber products are dense, the shipping methods offering the lowest cost per unit weight are preferred. Depending on destination, warehouse availability, and transportation conditions, some methods are preferred by certain buyers. In international trade, latex rubber is mostly shipped in 20-foot ocean containers. Inside the container, smaller containers are used to store the latex.[42]

Rubber shortage and global economics

There is growing concern for the future supply of rubber due to various factors, including plant disease, climate change, and the volatile market price of rubber.[43][44][45][46] Producers of natural rubber are mostly small family-held plantations, often serving large industrial aggregators. High volatility in the price of rubber affects rubber plantation investment, and farmers may remove their rubber trees if the international market spot price of a seemingly more profitable crop, (for example palm oil) surges in relation to rubber.

For instance, during the 2020 and 2021 international COVID-19 pandemic, demand for rubber gloves surged, leading to a spike in rubber prices of about 30%. In addition to the pandemic, demand exceeded supply in part because long term plantations had been torn out and replaced with other crops over the previous 5-10 years, and other areas were affected by climate-fueled natural disasters. In this environment, producers did increase their prices in keeping with supply and demand dynamics, putting upward price pressure on the whole downstream supply chain.[46]

Uses

Compression molded (cured) rubber boots before the flashes are removed

Uncured rubber is used for cements;[47] for adhesive, insulating, and friction tapes; and for crepe rubber used in insulating blankets and footwear. Vulcanized rubber has many more applications. Resistance to abrasion makes softer kinds of rubber valuable for the treads of vehicle tires and conveyor belts, and makes hard rubber valuable for pump housings and piping used in the handling of abrasive sludge.

The flexibility of rubber is appealing in hoses, tires and rollers for devices ranging from domestic clothes wringers to printing presses; its elasticity makes it suitable for various kinds of shock absorbers and for specialized machinery mountings designed to reduce vibration. Its relative gas impermeability makes it useful in the manufacture of articles such as air hoses, balloons, balls and cushions. The resistance of rubber to water and to the action of most fluid chemicals has led to its use in rainwear, diving gear, and chemical and medicinal tubing and as a lining for storage tanks, processing equipment and railroad tank cars. Because of their electrical resistance, soft rubber goods are used as insulation and for protective gloves, shoes, and blankets; hard rubber is used for articles such as telephone housings and parts for radio sets, meters, and other electrical instruments. The coefficient of friction of rubber, which is high on dry surfaces and low on wet surfaces, leads to its use for power-transmission belting, highly flexible couplings,[48] and for water-lubricated bearings in deep-well pumps. Indian rubber balls or lacrosse balls are made of rubber.

Compression molding machine for rubber parts

Around 25 million tonnes of rubber are produced each year, of which 30 percent is natural.[49] The remainder is synthetic rubber derived from petrochemical sources. The top end of latex production results in latex products such as surgeons' gloves, balloons, and other relatively high-value products. The mid-range which comes from the technically specified natural rubber materials ends up largely in tires but also in conveyor belts, marine products, windshield wipers, and miscellaneous goods. Natural rubber offers good elasticity, while synthetic materials tend to offer better resistance to environmental factors such as oils, temperature, chemicals, and ultraviolet light. "Cured rubber" is rubber that has been compounded and subjected to the vulcanisation process to create cross-links within the rubber matrix. Rubber can be added to cement to improve its properties.[50]

Allergic reactions

Some people have a serious latex allergy, and exposure to natural latex rubber products such as latex gloves can cause anaphylactic shock. The antigenic proteins found in Hevea latex are greatly reduced by about 99.9 percent (though not eliminated)[51] through vulcanization processing.

Latex from non-Hevea sources, such as guayule, can be used without allergic reaction by persons with an allergy to Hevea latex.[52]

Some allergic reactions are not to the latex itself, but from residues of chemicals used to accelerate the cross-linking process. Although this may be confused with an allergy to latex, it is distinct from it, typically taking the form of Type IV hypersensitivity in the presence of traces of specific processing chemicals.[51][53]

Microbial degradation

Natural rubber is susceptible to degradation by a wide range of bacteria.[54][55][56][57][58][59][60][61] The bacteria Streptomyces coelicolor, Pseudomonas citronellolis, and Nocardia spp. are capable of degrading vulcanized natural rubber.[62]

See also

References

Citations

  1. ^ a b Dunstan, Wyndham Rowland (1911). "Rubber" . In Chisholm, Hugh (ed.). Encyclopædia Britannica. Vol. 23 (11th ed.). Cambridge University Press. p. 795.
  2. ^ Sirimaporn Leepromrath, et al. "Rubber crop diversity and its influential factors in Thailand." Journal of Rubber Research 24.3 (2021): 461-473.
  3. ^ Muhammad Fadzli Ali, et al., "The dynamics of rubber production in Malaysia: Potential impacts, challenges and proposed interventions." Forest Policy and Economics 127 (2021): 102449.
  4. ^ Fadhlan Zuhdi, "The Indonesian natural rubber export competitiveness in global market." International Journal of Agriculture System 8.2 (2021): 130-139 online.
  5. ^ Legner, Erich Fred. "Rubber and Other Latex Products". University of California, Riverside.
  6. ^ Heim, Susanne (2002). Autarkie und Ostexpansion: Pflanzenzucht und Agrarforschung im Nationalsozialismus. Wallstein Verlag. ISBN 978-3-89244-496-1.
  7. ^ "Making Rubber from Dandelion Juice". Science Daily. 28 October 2013. Retrieved 22 November 2013.
  8. ^ Smith, James P. Jr/ (2006). "Plants & Civilization: An Introduction to the Interrelationships of Plants and People. Section 8.4, Latex Plants". Humboldt State University Botanical Studies Open Educational Resources and Data, Humboldt State University Digital Commons. pp. 137–141. Retrieved 8 June 2019.
  9. ^ Burns, Bill. "The Gutta Percha Company". History of the Atlantic Cable & Undersea Communications. Retrieved 14 February 2009.
  10. ^ a b Heinz-Hermann Greve "Rubber, 2. Natural" in Ullmann's Encyclopedia of Industrial Chemistry, 2000, Wiley-VCH, Weinheim. doi:10.1002/14356007.a23_225
  11. ^ Emory Dean Keoke, Kay Marie Porterfield. 2009. Encyclopedia of American Indian Contributions to the World: 15,000 Years of Inventions and Innovations. Infobase Publishing
  12. ^ Tully, John (2011). The Devil's Milk: A Social History of Rubber. NYU Press. ISBN 978-1-58367-260-0.
  13. ^ a b "Charles Marie de la Condamine". bouncing-balls.com.
  14. ^ Hosler, D.; Burkett, S.L.; Tarkanian, M.J. (1999). "Prehistoric polymers: Rubber processing in ancient Mesoamerica". Science. 284 (5422): 1988–1991. doi:10.1126/science.284.5422.1988. PMID 10373117.
  15. ^ Slack, Charles (2002). Noble obsession: Charles Goodyear, Thomas Hancock, and the race to unlock the greatest industrial secret of the nineteenth century. Hyperion. ISBN 978-0-7868-6789-9.
  16. ^ Jackson, Joe (2008). The Thief at the End of the World. Viking. ISBN 978-0-670-01853-6.
  17. ^ Goodman, Jordan (2009). The Devil and Mr Casement. Verso, 2009. p. 83. ISBN 978-1-84467-334-6. Retrieved 13 August 2023.
  18. ^ Casement, Roger (25 October 1997). The Amazon Journal of Roger Casement. Anaconda Editions. p. 85. ISBN 1-901990-00-1. Retrieved 13 August 2023.
  19. ^ "Natural rubber in India". Archived from the original on 1 October 2016.
  20. ^ Cornelius-Takahama, Vernon (2001). "Sir Henry Nicholas Ridley". Singapore Infopedia. Archived from the original on 4 May 2013. Retrieved 9 February 2013.
  21. ^ Leng, Dr Loh Wei; Keong, Khor Jin (19 September 2011). "Mad Ridley and the rubber boom". Malaysia History. Archived from the original on 27 July 2013. Retrieved 9 February 2013.
  22. ^ "Casing Joint Design" (PDF). Report – Investigation of the Challenger Accident. US Government Printing Office. Archived (PDF) from the original on 9 October 2022. Retrieved 29 August 2015.
  23. ^ Nur Fadhilah Idris; Nor Hidayaty Kamarulzaman; Zairossani Mohd Nor (2012). "Determination of Volatile Fatty Acids from Raw Natural Rubber Drying Activity by Thermal Desorption-Gas Chromatography" (PDF). Chemical Engineering Transactions. 30. doi:10.3303/CET1230030. S2CID 7469231. Archived from the original (PDF) on 15 December 2017. Retrieved 14 December 2017.
  24. ^ Hoven, Vipavee P.; Rattanakaran, Kesinee; Tanaka, Yasuyuki (1 November 2003). "Determination of Chemical Components that Cause Mal-Odor from Natural Rubber". Rubber Chemistry and Technology. 76 (5): 1128–1144. doi:10.5254/1.3547792.
  25. ^ a b c "Info" (PDF). aidic.it. Archived (PDF) from the original on 9 October 2022.
  26. ^ Koyama, Tanetoshi; Steinbüchel, Alexander, eds. (June 2011). "Biosynthesis of Natural Rubber and Other Natural Polyisoprenoids". Polyisoprenoids. Biopolymers. Vol. 2. Wiley-Blackwell. pp. 73–81. ISBN 978-3-527-30221-5.
  27. ^ Paterson-Jones, J.C.; Gilliland, M.G.; Van Staden, J. (June 1990). "The Biosynthesis of Natural Rubber". Journal of Plant Physiology. 136 (3): 257–263. doi:10.1016/S0176-1617(11)80047-7. ISSN 0176-1617.
  28. ^ Xie, W.; McMahan, C.M.; Distefano, A.J. DeGraw, M.D.; et al. (2008). "Initiation of rubber synthesis: In vitro comparisons of benzophenone-modified diphosphate analogues in three rubber producing species". Phytochemistry. 69 (14): 2539–2545. Bibcode:2008PChem..69.2539X. doi:10.1016/j.phytochem.2008.07.011. PMID 18799172.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  29. ^ Casey, P.J.; Seabra, M.C. (1996). "Protein Prenyltransferases". Journal of Biological Chemistry. 271 (10): 5289–5292. doi:10.1074/jbc.271.10.5289. PMID 8621375.
  30. ^ Kang, H.; Kang, M.Y.; Han, K.H. (2000). "Identification of Natural Rubber and Characterization of Biosynthetic Activity". Plant Physiol. 123 (3): 1133–1142. doi:10.1104/pp.123.3.1133. PMC 59076. PMID 10889262.
  31. ^ "Overview of the Causes of Natural Rubber Price Volatility". En.wlxrubber.com. 1 February 2010. Archived from the original on 26 May 2013. Retrieved 21 March 2013.
  32. ^ "Statistical Summary of World Rubber Situation" (PDF). International Rubber Study Group. December 2018. Archived (PDF) from the original on 5 February 2019. Retrieved 5 February 2019.
  33. ^ "Natural rubber leading producers worldwide 2021". Statista. Retrieved 18 February 2023.
  34. ^ Listiyorini, Eko (16 December 2010). "Rubber Exports From Indonesia May Grow 6%–8% Next Year". bloomberg.com. Archived from the original on 4 November 2012. Retrieved 21 March 2013.
  35. ^ Keoke, Emory (2003). Encyclopedia of American Indian contributions to the world 15,000 years of inventions and innovations. Checkmark Books. p. 156.
  36. ^ This section has been copied almost verbatim from the public domain UN Food and Agriculture Organization (FAO), ecoport.com article: Cecil, John; Mitchell, Peter; Diemer, Per; Griffee, Peter (2013). "Processing of Natural Rubber, Manufacture of Latex-Grade Crepe Rubber". ecoport.org. FAO, Agricultural and Food Engineering Technologies Service. Retrieved 19 March 2013.
  37. ^ Basic Rubber Testing. ASTM International. pp. 6–. GGKEY:8BT2U3TQN7G.
  38. ^ "Science of Plastics". Science History Institute. 18 July 2016. Retrieved 18 February 2023.
  39. ^ Schuster, Jens; Lutz, Johannes; Shaik, Yousuf Pasha; Yadavalli, Venkat Reddy (1 October 2022). "Recycling of fluoro-carbon-elastomers – A review". Advanced Industrial and Engineering Polymer Research. Recycling of Rubbers. 5 (4): 248–254. doi:10.1016/j.aiepr.2022.08.002. ISSN 2542-5048. S2CID 251658624.
  40. ^ Asaro, Lucia; Gratton, Michel; Seghar, Saïd; Aït Hocine, Nourredine (1 June 2018). "Recycling of rubber wastes by devulcanization". Resources, Conservation and Recycling. 133: 250–262. doi:10.1016/j.resconrec.2018.02.016. ISSN 0921-3449. S2CID 115671283.
  41. ^ Ducháček, Vratislav; Kuta, Antonín (October 1986). "Long-time sulfenamide-accelerated sulfur vulcanization of natural rubber/chlorobutyl rubber compounds". Journal of Applied Polymer Science. 32 (5): 4849–4855. doi:10.1002/app.1986.070320507. ISSN 0021-8995.
  42. ^ Transportation of Natural Rubber – Industry Source
  43. ^ Espinoza, Mauricio (1 November 2012). "US seeks domestic rubber as global shortage worsens". Ohio State University. Retrieved 13 March 2021. The importance of developing alternative sources of natural rubber becomes clear when considering that there has been a shortage in the supply of this critical resource every year since 2004. By 2020, the global shortfall of natural rubber is projected to be more than the entire amount (1.2 million metric tonnes) the U.S. imports every year.
  44. ^ "Natural rubber shortage". The Rubber Economist. Retrieved 13 March 2021.
  45. ^ Meyer, Robert (9 September 2019). "Situation Report: The looming crisis of natural rubber. The Chief Executive's view". Retrieved 13 March 2021.
  46. ^ a b Swain, Frank (8 March 2021). "The wonder material we all need but is running out: Climate change, capitalism and disease are threatening to strike a mortal blow to the world's rubber trees. Do we need to find alternative sources of rubber before it's too late?". BBC Future. Retrieved 13 March 2021.
  47. ^ Horath, Larry (2017). Fundamentals of Materials Science for Technologists: Properties, Testing, and Laboratory Exercises, Second Edition. Waveland Press. ISBN 978-1-4786-3518-5.
  48. ^ VULKAN Couplings System Competence – Compound Research, archived from the original on 11 December 2021, retrieved 9 June 2021
  49. ^ "Rubber Faqs". Archived from the original on 13 September 2016.
  50. ^ Liu, Luqing; Wang, Chaohui; Liang, Qing; Chen, Feng; Zhou, Xiaolei (15 March 2023). "A state-of-the-art review of rubber modified cement-based materials: Cement stabilized base". Journal of Cleaner Production. 392: 136270. doi:10.1016/j.jclepro.2023.136270. ISSN 0959-6526. S2CID 256564049.
  51. ^ a b "Premarket Notification [510(k)] Submissions for Testing for Skin Sensitization To Chemicals in Natural Rubber Products" (PDF). FDA. Retrieved 22 September 2013.
  52. ^ "New Type of Latex Glove Cleared". Food and Drug Administration.
  53. ^ American Latex Allergy Association. "Allergy Fact Sheet". Archived from the original on 13 March 2012. Retrieved 22 September 2013.
  54. ^ Rook, J.J. (1955). "Microbiological deterioration of vulcanized rubber". Appl. Microbiol. 3 (5): 302–309. doi:10.1128/aem.3.5.302-309.1955. PMC 1057125. PMID 13249390.
  55. ^ Leeang, K.W.H. (1963). "Microbiologic degradation of rubber". J. Am. Water Works Assoc. 53 (12): 1523–1535. Bibcode:1963JAWWA..55l1523L. doi:10.1002/j.1551-8833.1963.tb01176.x.
  56. ^ Tsuchii, A.; Suzuki, T.; Takeda, K. (1985). "Microbial degradation of natural rubber vulcanizates". Appl. Environ. Microbiol. 50 (4): 965–970. Bibcode:1985ApEnM..50..965T. doi:10.1128/AEM.50.4.965-970.1985. PMC 291777. PMID 16346923.
  57. ^ Heisey, R.M.; Papadatos, S. (1995). "Isolation of microorganisms able to metabolize puri¢ed natural rubber". Appl. Environ. Microbiol. 61 (8): 3092–3097. Bibcode:1995ApEnM..61.3092H. doi:10.1128/AEM.61.8.3092-3097.1995. PMC 1388560. PMID 16535106.
  58. ^ Jendrossek, D.; Tomasi, G.; Kroppenstedt, R.M. (1997). "Bacterial degradation of natural rubber: a privilege of actinomycetes?". FEMS Microbiology Letters. 150 (2): 179–188. doi:10.1016/s0378-1097(97)00072-4. PMID 9170260.
  59. ^ Linos, A. and Steinbuchel, A. (1998) Microbial degradation of natural and synthetic rubbers by novel bacteria belonging to the genus Gordona. Kautsch. Gummi Kunstst. 51, 496–499.
  60. ^ Linos, Alexandros; Steinbuchel, Alexander; Spröer, Cathrin; Kroppenstedt, Reiner M. (1999). "Gordonia polyisoprenivorans sp. nov., a rubber degrading actinomycete isolated from automobile tire". Int. J. Syst. Bacteriol. 49 (4): 1785–1791. doi:10.1099/00207713-49-4-1785. PMID 10555361.
  61. ^ Linos, Alexandros; Reichelt, Rudolf; Keller, Ulrike; Steinbuchel, Alexander (October 1999). "A Gram-negative bacterium, identified as Pseudomonas aeruginosa AL98, is a potent degrader of natural rubber and synthetic cis-1,4-polyisoprene". FEMS Microbiology Letters. 182 (1): 155–161. doi:10.1111/j.1574-6968.2000.tb08890.x. PMID 10612748.
  62. ^ Helge B. Bode; Axel Zeeck; Kirsten Plückhahn; Dieter Jendrossek (September 2000). "Physiological and Chemical Investigations into Microbial Degradation of Synthetic Poly(cis-1,4-isoprene)". Applied and Environmental Microbiology. 66 (9): 3680–3685. Bibcode:2000ApEnM..66.3680B. doi:10.1128/AEM.66.9.3680-3685.2000. PMC 92206. PMID 10966376.

Sources

  • Ali, Muhammad Fadzli, et al. "The dynamics of rubber production in Malaysia: Potential impacts, challenges and proposed interventions." Forest Policy and Economics 127 (2021): 102449.

Further reading

  • Dean, Warren. (1997) Brazil and the Struggle for Rubber: A Study in Environmental History. Cambridge University Press.
  • Grandin, Greg. Fordlandia: The Rise and Fall of Henry Ford's Forgotten Jungle City. Picador Press 2010. ISBN 978-0-312-42962-1
  • Weinstein, Barbara (1983) The Amazon Rubber Boom 1850–1920. Stanford University Press.
  • Tully, John A. The Devil's Milk; A Social History of Rubber. New York: Monthly Review Press, 2011.

External links

This page was last edited on 11 February 2024, at 06:29
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.