请输入您要查询的百科知识:

 

词条 Draft:Amphetamine synthesis
释义

  1. Routes of synthesis

      Via substitution at the 2-ethyl position    ...  

  2. References

  3. [Potentially relevant background material to be integrated, esp. for reaction attributions]

      Illegal synthesis  
{{AFC comment|1=Heavy use of primary literature citations and the big table under the section references induces confusion. The subject is not worth an article as contents are best covered here and here}} EvilxFish (talk) 08:19, 27 June 2018 (UTC){{AFC comment|1=Asked for opinions at CHEM. WBGconverse 04:56, 27 June 2018 (UTC)}}{{AFC comment|1=This is still clearly a work in progress but perhaps it should just be moved to mainspace to attract collaborators and avoid G13. Calliopejen1 (talk) 13:07, 7 June 2018 (UTC)}}

Syntheses of amphetamine and methamphetamine encompass a broad swath of organic chemistry, with more than seventy known chemical precursors. In some cases a wide variety of reagents have been applied to bring about the same reaction. Funded by pharmaceutical companies, government grants, and illegal chemists, the reactions have progressed in step with organic chemistry as a whole.[1][2] Exploration of a broader range of designer drugs based on the phenethylamine skeleton led to the syntheses of potent amphetamine derivatives with very different psychoactive effects, such as MDMA ("Ecstasy") and trimethoxyamphetamine, a mescaline-like compound.[3]

Routes of synthesis

The following reaction schemes are organized according to retrosynthetic analysis, i.e. the most immediate precursor is described, followed by an explanation of the reaction schemes used to produce it.

Via substitution at the 2-ethyl position

Many popular syntheses involve removing some functional group from the 2-ethyl position of precursor compounds which may be natural products or legally obtainable. Commonly this is a hydroxyl group, found in ephedrine and pseudoephedrine (which differ only in the orientation of the -OH).[4][5] The (pseudo)ephedrine may in turn result from reduction of another compound, such as an oxime with a ketone at the 2-position.[5]

Nagai Nagayoshi first isolated ephedrine from the Chinese medicinal herb Ephedra (Ma Huang). In 1893 he published a synthesis of methamphetamine by the removal of the hydroxyl.[4][8][5] In 1919, pharmacologist Akira Ogata published a synthesis of methamphetamine hydrochloride via reduction of ephedrine using red phosphorus and iodine.[5] Though dating back to the discovery of the drug, the Nagai route[6] did not become popular among illicit manufacturers until ca. 1982.[7] Either hydrogen iodide or iodine and water may be used with red phosphorus in this reaction.[8][14] On heating the precursor is rapidly iodinated by the hydrogen iodide to form iodoephedrine. The phosphorus assists in the second step, by consuming iodine to form phosphorus triiodide (which decomposes in water to phosphorous acid, regenerating hydrogen iodide). Because hydrogen iodide exists in a chemical equilibrium with iodine and hydrogen, the phosphorus reaction shifts the balance toward hydrogen production when iodine is consumed (see Le Châtelier's principle).[9] In Australia, criminal groups have been known to substitute "red" phosphorus with either hypophosphorous acid or phosphorous acid (the "Hypo route").[8][10][11] This is a hazardous process for amateur chemists because phosphine gas, a side-product from in situ hydrogen iodide production,[12] is extremely toxic to inhale. The reaction can also create toxic, flammable white phosphorus waste.[14] Methamphetamine produced in this way is usually more than 95% pure.[13]

The conceptually similar Emde route involves reduction of ephedrine to chloroephedrine using thionyl chloride (SOCl2), followed by catalytic hydrogenation. The catalysts for this reaction are palladium or platinum.[22][14] The Rosenmund route also uses hydrogen gas and a palladium catalyst poisoned with barium sulfate (Rosenmund reduction), but uses perchloric acid instead of thionyl chloride.[15]

The Birch reduction, also called the "Nazi method", became popular in the mid-to-late 1990s and comprised the bulk of methamphetamine production in Michigan in 2002.[7] It reacts pseudoephedrine with liquid anhydrous ammonia and an alkali metal such as sodium or lithium. The reaction is allowed to stand until the ammonia evaporates.[16] However, the Birch reduction is dangerous because the alkali metal and ammonia are both extremely reactive, and the temperature of liquid ammonia makes it susceptible to explosive boiling when reactants are added. It has been the most popular method in Midwestern states of the U. S. because of the ready availability of liquid ammonia fertilizer in farming regions.[17][18]

In recent years, a simplified "Shake 'n Bake" one-pot synthesis has become more popular. The method is suitable for such small batches that pseudoephedrine restrictions are less effective, it uses chemicals that are easier to obtain (though no less dangerous than traditional methods), and it is so easy to carry out that some addicts have made the drug while driving.[19] It involves placing crushed pseudoephedrine tablets into a nonpressurized container containing ammonium nitrate, water, and a hydrophobic solvent such as Coleman fuel[20] or automotive starting fluid, to which lye and lithium (from lithium batteries) is added. Hydrogen chloride gas produced by a reaction of salt with sulfuric acid is then used to recover crystals for purification. The container needs to be "burped" periodically to prevent failure under accumulating pressure, as exposure of the lithium to the air can spark a flash fire.[20] The battery lithium can react with water to shatter a container and potentially start a fire or explosion.[20]

...

[1-phenylpropan-2-imine]

Achiral but reacts with Ir-(S,S)-f-binaphane and H2 to produce chiral dextroamphetamine

From 2-phenylacetonitrile via MeMgI[21][22]

E-(2-nitroprop-1-enyl)benzene

Reacts with 1-phenylpropan-3-amine and SmI2 to produce racemic amphetamine[23][24]

[*** The following should be listed under the azide, once I confirm with the original reference]

S-1-phenylpropan-2-ol

This is reacted with sodium azide, followed by reduction with H2 on Pd/C catalyst.[25][26]

From 2-benzyloxirane via LiAlH4, which is from (R)-3-phenylpropane-1,2-diol via TsCl and sodium hydride, which is from (R)-3-phenyl-2-(phenylaminooxy)propan-1-ol via Pd/C, which is produced from 3-phenylpropanal with nitrosobenzene, L-proline, and sodium borohydride.

References

CitationDateAvery and Cantrell
1989 ref #
Avery and Eli
2009 ref #
Malek[27]198846
Vogel[28]198764
Carey and Sundberg[29]197787
House[30]197290
Vogel[31]194875, 77
Zhong et al.[32]2005-09-289
Boswell and Lo[33]2002-06-0423
Ito and Nemori[34]199129
Stirling et al.[35]1990-08-2143
Ishiyama et al.[36]198848
Barfknecht and Nichols[37]197650
Buenger et al.[38]200953
Saigo et al.[39]199966
Bryan[40]1969-07-291370
Green[41]1965-06-012975
Metzger[42]1958-03-06377
Gillingham[43]196278
Tindall[44]1958-03-252579
Temmler[45]195323, 4583
Tindall[46]1953-08-0486
Stochdorph and von Schickh[47]1952-09-013387
[48]1954-01-27995
Tindall[49]1953-04-283296
Flisik et al.[50]1946105
Zienty[51]195881b
Temmler[52]1939-07-2015, 22
Dobke and Keil[53]1952-01-3118
Nakajima[54]1951-05-1519
Dobke and Keil[55]195222
Laboratoires Amido[56]1964-10-0540
Keil and Dobke[57]194041
Keil and Dobke[58]1953-03-1941
Temmler[59]1939-07-2043
Alles[60]1932-0950113
Heinzelman[61]195153
Shinohara et al.[62]196462
Dobke and Keil[63]1938-10-036, 18
Hou et al.[64]20091
Routaboul et al.[65]20082
Nechab et al.[66]20073
Ankner and Hilmersson[67]20074
Talluri and Sudalai[68]20075
Granander et al.[69]20066
Guy et al.[70]20087
Barrett et al.[71]20058
Yang et al.[72]200510
Fecik et al.[73]200511
Kinderman et al.[74]200512
Guisado et al.[75]200513
Shi et al.[76]200414
Kitamura et al.[77]200415
Sayyed and Sudalai[78]200416
Mukade et al.[79]200317
Li and Izumi[80]200318
Wagner et al.[81]200319
Li and Izumi[82]200220
Davies and Dixon[83]200221
Gonzales-Sabin et al.[84]200222
Davis and Durden[85]200124
Quagliato et al.[86]200025
Haak et al.[87]200026
Hartung et al.[88]200027
Subramanyam[89]200028
Liu et al.[90]199630
Gajda et al.[91]199731
IMicovic et al.[92]199632
Donner[93]199533
Rozwadowska[94]199334
Schoeps and Halldin[95]199235
Krasik and Alper[96]199236
Gee and Langstrom[97]199137
Acs et al.[98]199138
Mori et al.[99]199039
Noggle, Jr. et al.[100]199040
Harsy[101]199041
Kabalka et al.[102]199042
Giannis and Sandhoff[103]198944
Jilek et al.[104]198945
Smith et al.[105]198847
Denmark et al.[106]198749
Brunner et al.[107]198651
Baldwin et al.[108]198652
Grishina et al.[109]198554
Nordlander et al.[110]198555
Varma and Kabalka[111]198556
Mourad et al.[112]198457
Koziara et al.[113]198258
Krzyzanowska and Stec[114]198259
Finn et al.[115]198160
Tuaillon and Perrot[116]197861
Beckett et al.[117]197662
Wassink et al.[118]19744463
Nichols et al.[119]197364
Borch et al.[120]19715765
Ho et al.[121]197067
Foreman et al.[122]196968
Brown and Kurek[123]19695669
Cervinka et al.[124]19685871
Milstein[125]196872
Kotera et al.[126]19686073
Iwai et al.[127]196574
Lukasiewicz[128]19637276
Schrecker[129]195780
Kametani and Nomura[130]19543182
Gilsdorf and Nord[131]19526185
Acs et al.[132]199488
Max and Ramirez[133]195089
Ingersoll[134]193790
Wilson III[135]195030, 8091
Mastigle et al.[136]19502692
Wessling and Schaefer[137]199193
Hass et al.[138]195035, 5294
Alexander and Wildman[139]19487197
Alexander and Misegades[140]19482098
Kametani and Nomura[141]195499
Haskelberg[142]194827100
Muthukumaran and Krishnan[143]1992101
JRitter and Kalish[144]1948102
Kindler et al.[145]194814, 17103
Patrick Jr. et al.[146]1946104
Crossley and Moore[147]194465106
Bobranskii and Drabik[148]194168, 79107
Magidson and Garkuska[149]194167, 73108
Lane et al.[150]1940109
Braun and Friehmelt[151]1933110
Wallis and Dripps[152]1933111
Gordon[153]1932112
Leithe[154]193237, 49114
Hartung and Munch[155]193111115
Wallis and Nagel[156]1933116
Hey[157]193051117
Hartung[158]192812118
Potapov and Terent‘ev[159]195881a
Gal[160]197784a
Repke et al.[161]19785984b
Sinnema and Verweji[162]19811
Hider[163]19692
Emde[164]19295
Larizza et al.[165]19667
Rosenmund et al.[166]19428
Hartung and Chang[167]195216
Manske and Johnson[168]192924
Novelli[169]193928
Fusco and Canonica[170]195034
Cook[171]196236
Evdokimoff[172]195138
Ogata[173]191946
Erlenmeyer and Simon[174]194247
Jaeger and van Dijk[175]194148
Weichet et al.[176]196155
Cervinka et al.[177]196866
Ogata[178]191969
Moore[179]194970
Herbst and Manske[180]193676
Julian et al.[181]198478
Walker and Hauser[182]194681
Sund and Henze[183]197082
McKillop and Hunt[184]197083
Tiffeneau[185]190784
Mason and Terry[186]194085
King and McMillan[187]195186
Lebrilla and Maier[188]198688
Maier et al.[189]198689
Cooper[190]198591
Noggle Jr. et al.[191]198692
Allen and Kiser[192]198793
Tsutsumi[193]195310, 74
Shiho[194]194421, 42
Gero[195]19514, 39, 54
Ho and Wong[196]197563(a)
Parham and Sayed[197]197663(b)
Shaw et al.[198]195663(c)
Cope et al.[199]196263(d)
Reusch and LeMahiem[200]196463(e)
Marvel et al.[201]194163(f)
Chen et al.[202]2009

[Potentially relevant background material to be integrated, esp. for reaction attributions]

{{main|History and culture of substituted amphetamines}}Amphetamine was first synthesized in 1887 in Germany by Romanian chemist Lazăr Edeleanu who named it phenylisopropylamine.[203][204][205] Shortly after, methamphetamine was synthesized from ephedrine in 1893 by Japanese chemist Nagai Nagayoshi.[206]

Amphetamine was first synthesized in 1887 in Germany by Romanian chemist, Lazăr Edeleanu, who named the drug phenylisopropylamine.[205][207][208][209]

Methamphetamine constituted half of the amphetamine salts for the original formulation for the diet drug Obetrol.[5]

In 1997 and 1998, researchers at Texas A&M University claimed to have found amphetamine and methamphetamine in the foliage of two Acacia species native to Texas, A. berlandieri and A. rigidula.[212][213] Previously, both of these compounds had been thought to be purely synthetic.[212][213][233] These findings have never been duplicated and consequently the validity of the report has come into question.[214]

As early as 1919, Akira Ogata synthesized methamphetamine via reduction of ephedrine using red phosphorus and iodine. Later, the chemists Hauschild and Dobke from the German pharmaceutical company Temmler developed an easier method for converting ephedrine to methamphetamine. As a result, it was possible for Temmler to market it on a large scale as a nonprescription drug under the trade name Pervitin (methamphetamine hydrochloride). It was not until 1986 that Pervitin became a controlled substance, requiring a special prescription to obtain.[215] Pervitin was commonly used by the German and Finnish militaries.[216][217]

Illegal synthesis

Methamphetamine is most structurally similar to methcathinone and amphetamine. Synthesis is relatively simple, but entails risk with flammable and corrosive chemicals, particularly the solvents used in extraction and purification. The six major routes of production begin with either phenyl-2-propanone (P2P) or with one of the isomeric compounds pseudoephedrine and ephedrine.[218]

One procedure uses the reductive amination of phenylacetone with methylamine,[219] P2P was usually obtained from phenylacetic acid and acetic anhydride,[220] and phenylacetic acid might arise from benzaldehyde, benzylcyanide, or benzylchloride.[15] Methylamine is crucial to all such methods, and is produced from the model airplane fuel nitromethane, or formaldehyde and ammonium chloride, or methyl iodide with hexamine.[221] This was once the preferred method of production by motorcycle gangs in California,[222] until DEA restrictions on the chemicals made the process difficult. Pseudoephedrine, ephedrine, phenylacetone, and phenylacetic acid are currently DEA list I and acetic anhydride is list II on the DEA list of chemicals subject to regulation and control measures. This method can involve the use of mercuric chloride and leaves behind mercury and lead environmental wastes.[17] The methamphetamine produced by this method is racemic, consisting partly of the less-desired levomethamphetamine isomer,[223] though separation of the two enantiomeric forms through selective recrystallization of the tartrate salt can occur in order to isolate the more active dextromethamphetamine.[224]

The alternative Leuckart route also relies on P2P to produce a racemic product, but proceeds via methylformamide in formic acid to an intermediate N-formyl-methamphetamine, which is then decarboxylated with hydrochloric acid.[218][15]

Illicit methamphetamine is more commonly made by the reduction of ephedrine or pseudoephedrine, which produces the more active d-methamphetamine isomer. The maximum conversion rate for ephedrine and pseudoephedrine is 92%, although typically, illicit methamphetamine laboratories convert at a rate of 50% to 75%.[225] Most methods of illicit production involve protonation of the hydroxyl group on the ephedrine or pseudoephedrine molecule.

Though dating back to the discovery of the drug, the Nagai route[226] did not become popular among illicit manufacturers until ca. 1982, and comprised 20% of production in Michigan in 2002.[7] It involves red phosphorus and hydrogen iodide (also known as hydroiodic acid or iohydroic acid). (The hydrogen iodide is replaced by iodine and water in the "Moscow route"[8]) The hydrogen iodide is used to reduce either ephedrine or pseudoephedrine to methamphetamine.[17] On heating the precursor is rapidly iodinated by the hydrogen iodide to form iodoephedrine. The phosphorus assists in the second step, by consuming iodine to form phosphorus triiodide (which decomposes in water to phosphorous acid, regenerating hydrogen iodide). Because hydrogen iodide exists in a chemical equilibrium with iodine and hydrogen, the phosphorus reaction shifts the balance toward hydrogen production when iodine is consumed (see Le Châtelier's principle).[9] In Australia, criminal groups have been known to substitute "red" phosphorus with either hypophosphorous acid or phosphorous acid (the "Hypo route").[8][10][11] This is a hazardous process for amateur chemists because phosphine gas, a side-product from in situ hydrogen iodide production,[12] is extremely toxic to inhale. The reaction can also create toxic, flammable white phosphorus waste.[17] Methamphetamine produced in this way is usually more than 95% pure.[13]

The conceptually similar Emde route involves reduction of ephedrine to chloroephedrine using thionyl chloride (SOCl2), followed by catalytic hydrogenation. The catalysts for this reaction are palladium or platinum.[218][14] The Rosenmund route also uses hydrogen gas and a palladium catalyst poisoned with barium sulfate (Rosenmund reduction), but uses perchloric acid instead of thionyl chloride.[15]

The Birch reduction, also called the "Nazi method", became popular in the mid-to-late 1990s and comprised the bulk of methamphetamine production in Michigan in 2002.[7] It reacts pseudoephedrine with liquid anhydrous ammonia and an alkali metal such as sodium or lithium. The reaction is allowed to stand until the ammonia evaporates.[16] However, the Birch reduction is dangerous because the alkali metal and ammonia are both extremely reactive, and the temperature of liquid ammonia makes it susceptible to explosive boiling when reactants are added. It has been the most popular method in Midwestern states of the U. S. because of the ready availability of liquid ammonia fertilizer in farming regions.[17][18]

In recent years, a simplified "Shake 'n Bake" one-pot synthesis has become more popular. The method is suitable for such small batches that pseudoephedrine restrictions are less effective, it uses chemicals that are easier to obtain (though no less dangerous than traditional methods), and it is so easy to carry out that some addicts have made the drug while driving.[19] It involves placing crushed pseudoephedrine tablets into a nonpressurized container containing ammonium nitrate, water, and a hydrophobic solvent such as Coleman fuel[20] or automotive starting fluid, to which lye and lithium (from lithium batteries) is added. Hydrogen chloride gas produced by a reaction of salt with sulfuric acid is then used to recover crystals for purification. The container needs to be "burped" periodically to prevent failure under accumulating pressure, as exposure of the lithium to the air can spark a flash fire.[20] The battery lithium can react with water to shatter a container and potentially start a fire or explosion.[20]

Rarely, the impure reaction mixture from the hydrogen iodide/red phosphorus route is used without further modification, usually by injection; it is called "ox blood".[16] "Meth oil" refers to the crude methamphetamine base produced by several synthesis procedures. Ordinarily it is purified by exposure to hydrogen chloride, as a solution or as a bubbled gas, and extraction of the resulting salt occurs by precipitation and/or recrystallization with ether/acetone.[16]

[227]
随便看

 

开放百科全书收录14589846条英语、德语、日语等多语种百科知识,基本涵盖了大多数领域的百科知识,是一部内容自由、开放的电子版国际百科全书。

 

Copyright © 2023 OENC.NET All Rights Reserved
京ICP备2021023879号 更新时间:2024/9/20 15:01:35