Хидролиза — разлика између измена

С Википедије, слободне енциклопедије
Садржај обрисан Садржај додат
м Бот: исправљена преусмерења
.
ознака: везе до вишезначних одредница
Ред 1: Ред 1:
{{Short description|Цепање хемијских веза додавањем воде}}
[[Датотека:Hydrolysis.png|мини|350п|десно|Генерички механизам за хидролошку реакцију]]
[[Датотека:Hydrolysis.png|мини|350п|десно|Генерички механизам за хидролошку реакцију]]

'''Хидролиза''' је [[хемијска реакција]] која се заснива на '''распаду [[молекул]]а [[хемијско једињење|хемијских једињења]]''' на два мања фрагмента под утицајем контакта са [[вода|водом]] или воденом паром. У посебном случају [[хидролиза соли|хидролизе соли]] то је реакција [[јон]]а са водом (углавном се ствара кисела или базна средина). Хидролиза је посебна грана [[солволиза|солволизе]].
'''Хидролиза''' је [[хемијска реакција]] која се заснива на ''распаду [[молекул]]а [[хемијско једињење|хемијских једињења]]'' на два мања фрагмента под утицајем контакта са [[вода|водом]] или воденом паром. The term is used broadly for [[substitution reaction|substitution]], [[elimination reaction|elimination]], and [[solvation]] reactions in which water is the [[nucleophile]].<ref>{{GoldBookRef | title = Hydrolysis | file = H02902}}{{GoldBookRef | title = Solvolysis| file = S05762}}</ref> У посебном случају [[хидролиза соли|хидролизе соли]], то је реакција [[јон]]а са водом (углавном се ствара кисела или базна средина). Хидролиза је посебна грана [[солволиза|солволизе]].
{{рут}}
Biological hydrolysis is the cleavage of biomolecules where a water molecule is consumed to effect the separation of a larger molecule into component parts. When a [[carbohydrate]] is broken into its component sugar molecules by hydrolysis (e.g., [[sucrose]] being broken down into [[glucose]] and [[fructose]]), this is recognized as [[saccharification]].<ref name=saccharification>{{cite web |title=Definition of Saccharification |url=https://www.merriam-webster.com/dictionary/saccharification |website=www.merriam-webster.com |access-date=8 September 2020 |language=en |archive-date=7 January 2021 |archive-url=https://web.archive.org/web/20210107190250/https://www.merriam-webster.com/dictionary/saccharification |url-status=live }}</ref>

Hydrolysis reactions can be the reverse of a [[condensation reaction]] in which two molecules join into a larger one and eject a water molecule. Thus hydrolysis adds water to break down, whereas condensation builds up by removing water.<ref>{{cite web |last1=Steane |first1=Richard |title=Condensation and Hydrolysis |url=https://www.biotopics.co.uk/as/condensation_and_hydrolysis.html |website=www.biotopics.co.uk |access-date=2020-11-13 |archive-date=2020-11-27 |archive-url=https://web.archive.org/web/20201127013953/https://biotopics.co.uk/as/condensation_and_hydrolysis.html |url-status=live }}</ref>

== Типови ==
Usually hydrolysis is a chemical process in which a molecule of water is added to a substance. Sometimes this addition causes both substance and water molecule to split into two parts. In such reactions, one fragment of the target molecule (or parent molecule) gains a [[hydrogen ion]]. It breaks a chemical bond in the compound.

===Salts===
A common kind of hydrolysis occurs when a [[salt (chemistry)|salt]] of a [[weak acid]] or [[weak base]] (or both) is dissolved in water. [[Self-ionization of water|Water spontaneously ionizes]] into [[Hydroxide|hydroxide anions]] and [[Hydronium|hydronium cations]]. The salt also dissociates into its constituent anions and cations. For example, [[sodium acetate]] dissociates in water into [[sodium]] and [[acetate]] ions. Sodium ions react very little with the hydroxide ions whereas the acetate ions combine with hydronium ions to produce [[acetic acid]]. In this case the net result is a relative excess of hydroxide ions, yielding a basic [[Solution (chemistry)|solution]].

[[Strong acids]] also undergo hydrolysis. For example, dissolving [[sulfuric acid]] (H<sub>2</sub>SO<sub>4</sub>) in water is accompanied by hydrolysis to give [[hydronium]] and [[bisulfate]], the sulfuric acid's [[conjugate acid|conjugate base]]. For a more technical discussion of what occurs during such a hydrolysis, see [[Brønsted–Lowry acid–base theory]].

===Esters and amides===
Acid–base-catalysed hydrolyses are very common; one example is the hydrolysis of [[amide]]s or [[ester]]s. Their hydrolysis occurs when the [[nucleophile]] (a nucleus-seeking agent, e.g., water or hydroxyl ion) attacks the carbon of the [[carbonyl|carbonyl group]] of the [[ester]] or [[amide]]. In an aqueous base, hydroxyl ions are better nucleophiles than polar molecules such as water. In acids, the carbonyl group becomes protonated, and this leads to a much easier nucleophilic attack. The products for both hydrolyses are compounds with [[carboxylic acid]] groups.

Perhaps the oldest commercially practiced example of ester hydrolysis is [[saponification]] (formation of soap). It is the hydrolysis of a [[triglyceride]] (fat) with an aqueous base such as [[sodium hydroxide]] (NaOH). During the process, [[glycerol]] is formed, and the [[fatty acid]]s react with the base, converting them to salts. These salts are called soaps, commonly used in households.

In addition, in living systems, most biochemical reactions (including ATP hydrolysis) take place during the catalysis of [[enzyme]]s. The catalytic action of enzymes allows the hydrolysis of [[protein]]s, fats, oils, and [[carbohydrate]]s. As an example, one may consider [[protease]]s (enzymes that aid [[digestion]] by causing hydrolysis of [[peptide bond]]s in [[protein]]s). They catalyze the hydrolysis of interior peptide bonds in peptide chains, as opposed to [[exopeptidase]]s (another class of enzymes, that catalyze the hydrolysis of terminal peptide bonds, liberating one free amino acid at a time).

However, proteases do not catalyze the hydrolysis of all kinds of proteins. Their action is stereo-selective: Only proteins with a certain tertiary structure are targeted as some kind of orienting force is needed to place the amide group in the proper position for catalysis. The necessary contacts between an enzyme and its substrates (proteins) are created because the enzyme folds in such a way as to form a crevice into which the substrate fits; the crevice also contains the catalytic groups. Therefore, proteins that do not fit into the crevice will not undergo hydrolysis. This specificity preserves the integrity of other proteins such as [[hormone]]s, and therefore the biological system continues to function normally.
[[File:Acid-CatAmideHydrolMarch.png|thumb|upright=1.7|Mechanism for acid-catalyzed hydrolysis of an amide.]]

Upon hydrolysis, an [[amide]] converts into a [[carboxylic acid]] and an [[amine]] or [[ammonia]] (which in the presence of acid are immediately converted to ammonium salts). One of the two oxygen groups on the carboxylic acid are derived from a water molecule and the amine (or ammonia) gains the hydrogen ion. The hydrolysis of [[peptide bond|peptides]] gives [[amino acid]]s.

Many [[polyamide]] polymers such as [[nylon 6,6]] hydrolyze in the presence of strong acids. The process leads to [[depolymerization]]. For this reason nylon products fail by fracturing when exposed to small amounts of acidic water. Polyesters are also susceptible to similar [[polymer degradation]] reactions. The problem is known as [[environmental stress cracking]].

===ATP===
Hydrolysis is related to [[energy metabolism]] and storage.<ref> Nelson, David L., Cox, Michael M. ''Lehninger: Principles of Biochemistry''. New York: W.H. Freeman and Company, 2013. Sixth ed., p. 522–523. </ref><ref>Hardie, D. G., Ross, F. A., Hawley, S. A (2012). [http://www.nature.com/nrm/journal/v13/n4/pdf/nrm3311.pdf AMPK: a nutrient and energy sensor that maintains energy homeostasis]. ''Nature'', '''13''', 251–262. Accessed 9 April 2017.</ref> All living cells require a continual supply of energy for two main purposes: the [[biosynthesis]] of micro and macromolecules, and the active transport of ions and molecules across cell membranes. The energy derived from the [[oxidation]] of nutrients is not used directly but, by means of a complex and long sequence of reactions, it is channeled into a special energy-storage molecule, [[adenosine triphosphate]] (ATP). The ATP molecule contains [[pyrophosphate]] linkages (bonds formed when two phosphate units are combined) that release energy when needed. ATP can undergo hydrolysis in two ways: Firstly, the removal of terminal phosphate to form [[adenosine diphosphate]] (ADP) and inorganic phosphate, with the reaction:

: ATP + {{chem|H|2|O}} → ADP + P<sub>i</sub>

Secondly, the removal of a terminal diphosphate to yield [[adenosine monophosphate]] (AMP) and [[pyrophosphate]]. The latter usually undergoes further cleavage into its two constituent phosphates. This results in biosynthesis reactions, which usually occur in chains, that can be driven in the direction of synthesis when the phosphate bonds have undergone hydrolysis.

=== Полисахариди {{Anchor|saccharification}}===
[[File:Sucrose-inkscape.svg|thumb|[[Сахароза]].<ref name=Sherris>{{cite book | veditors = Ryan KJ, Ray CG | title = Sherris Medical Microbiology | edition = 4th | publisher = McGraw Hill | year = 2004 | isbn= 978-0-8385-8529-0 }}</ref> The glycoside bond is represented by the central oxygen atom, which holds the two monosaccharide units together.]]
[[Monosaccharide]]s can be linked together by [[glycosidic bond]]s, which can be cleaved by hydrolysis. Two, three, several or many monosaccharides thus linked form [[disaccharide]]s, [[trisaccharide]]s, [[oligosaccharide]]s, or [[polysaccharide]]s, respectively. Enzymes that hydrolyze glycosidic bonds are called "[[glycoside hydrolase]]s" or "glycosidases".

The best-known disaccharide is [[sucrose]] (table sugar). Hydrolysis of sucrose yields [[glucose]] and [[fructose]]. [[Invertase]] is a [[sucrase]] used industrially for the hydrolysis of sucrose to so-called [[invert sugar]]. [[Lactase]] is essential for digestive hydrolysis of [[lactose]] in milk; many adult humans do not produce lactase and [[lactose intolerance|cannot digest the lactose]] in milk.

The hydrolysis of polysaccharides to soluble sugars can be recognized as [[saccharification]].<ref name=saccharification/> Malt made from [[barley]] is used as a source of β-amylase to break down [[starch]] into the disaccharide [[maltose]], which can be used by yeast to [[brewing|produce beer]]. Other [[amylase]] enzymes may convert starch to glucose or to oligosaccharides. [[Cellulose]] is first hydrolyzed to [[cellobiose]] by [[cellulase]] and then cellobiose is further hydrolyzed to [[glucose]] by [[beta-glucosidase]]. [[Ruminants]] such as cows are able to hydrolyze cellulose into cellobiose and then glucose because of [[symbiosis|symbiotic]] bacteria that produce cellulases.

===Metal aqua ions===
{{Main|Metal ions in aqueous solution}}

Metal ions are [[Lewis acid]]s, and in [[aqueous solution]] they form [[metal aquo complex]]es of the general formula M(H<sub>2</sub>O)<sub>n</sub><sup>m+</sup>.<ref>{{cite book |last1=Burgess |first1=John |title=Metal Ions in Solution |date=1978 |publisher=Ellis Horwood |location=Chichester |isbn=978-0853120278}}</ref><ref>{{cite book |last= Richens |first= D. T. |title= The Chemistry of Aqua Ions: Synthesis, Structure, and Reactivity: A Tour through the Periodic Table of the Elements |publisher= Wiley |year= 1997 |isbn= 0-471-97058-1}}</ref> The aqua ions undergo hydrolysis, to a greater or lesser extent. The first hydrolysis step is given generically as

:M(H<sub>2</sub>O)<sub>n</sub><sup>m+</sup> + H<sub>2</sub>O {{eqm}} M(H<sub>2</sub>O)<sub>n−1</sub>(OH)<sup>(m−1)+</sup> + H<sub>3</sub>O<sup>+</sup>

Thus the aqua [[cation]]s behave as acids in terms of [[Brønsted–Lowry acid–base theory]].<ref>{{cite journal | author-link = Johannes Nicolaus Brønsted |author=Brönsted, J. N. |year=1923|title=Einige Bemerkungen über den Begriff der Säuren und Basen|trans-title=Some observations about the concept of [[acids and bases]]|journal=Recueil des Travaux Chimiques des Pays-Bas|volume=42|issue=8|pages= 718–728|doi=10.1002/recl.19230420815}}</ref><ref>{{cite journal|author-link = Martin Lowry|author=Lowry, T. M. |year=1923|title=The uniqueness of hydrogen|doi=10.1002/jctb.5000420302|url=https://archive.org/stream/ost-chemistry-chemistryindustr01soci/chemistryindustr01soci#page/n65/mode/2up |journal=Journal of the Society of Chemical Industry|volume=42|issue=3|pages=43–47}}</ref> This effect is easily explained by considering the [[inductive effect]] of the positively charged metal ion, which weakens the O-H bond of an attached water molecule, making the liberation of a proton relatively easy.

The [[acid dissociation constant|dissociation constant]], pK<sub>a</sub>, for this reaction is more or less linearly related to the charge-to-size ratio of the metal ion.<ref name="bm">{{cite book |last1=Baes |first1=Charles F. |last2=Mesmer |first2=Robert E. |title=The Hydrolysis of Cations |date=1976 |publisher=Wiley |location=New York |isbn=9780471039853}}</ref> Ions with low charges, such as Na<sup>+</sup> are very weak acids with almost imperceptible hydrolysis. Large divalent ions such as Ca<sup>2+</sup>, Zn<sup>2+</sup>, Sn<sup>2+</sup> and Pb<sup>2+</sup> have a pK<sub>a</sub> of 6 or more and would not normally be classed as acids, but small divalent ions such as Be<sup>2+</sup> undergo extensive hydrolysis. Trivalent ions like Al<sup>3+</sup> and Fe<sup>3+</sup> are weak acids whose pK<sub>a</sub> is comparable to that of [[acetic acid]]. Solutions of salts such as BeCl<sub>2</sub> or Al(NO<sub>3</sub>)<sub>3</sub> in water are noticeably [[acidic]]; the hydrolysis can be [[Le chatelier's principle|suppressed]] by adding an acid such as [[nitric acid]], making the solution more acidic.

Hydrolysis may proceed beyond the first step, often with the formation of polynuclear species via the process of [[olation]].<ref name="bm"/> Some "exotic" species such as Sn<sub>3</sub>(OH)<sub>4</sub><sup>2+</sup><ref>{{Greenwood&Earnshaw|page=384}}</ref> are well characterized. Hydrolysis tends to proceed as [[pH]] rises leading, in many cases, to the precipitation of a hydroxide such as Al(OH)<sub>3</sub> or AlO(OH). These substances, major constituents of [[bauxite]], are known as [[laterite]]s and are formed by leaching from rocks of most of the ions other than aluminium and iron and subsequent hydrolysis of the remaining aluminium and iron.


== Реакције хидролизе ==
== Реакције хидролизе ==
Ред 33: Ред 89:


== Референце ==
== Референце ==
{{reflist}}
{{reflist|}}

== Литература ==
{{reflist|}}
* {{cite journal |doi=10.1021/ar000209h|title=Catalytic Functionalization of Arenes and Alkanes via C−H Bond Activation|year=2001|last1=Jia|first1=Chengguo|last2=Kitamura|first2=Tsugio|last3=Fujiwara|first3=Yuzo|journal=Accounts of Chemical Research|volume=34|issue=8|pages=633–639|pmid=11513570}}
* {{cite journal |doi=10.1021/cr00032a009|title=Catalytic Asymmetric Dihydroxylation|year=1994|last1=Kolb|first1=Hartmuth C.|last2=Vannieuwenhze|first2=Michael S.|last3=Sharpless|first3=K. Barry|journal=Chemical Reviews|volume=94|issue=8|pages=2483–2547}}
* {{cite journal |author=Huang, X. |author2=Groves, J. T. |author-link2=John T. Groves |title=Beyond ferryl-mediated hydroxylation: 40 years of the rebound mechanism and C–H activation |journal=JBIC Journal of Biological Inorganic Chemistry |year=2017 |volume=22 |issue=2–3 |pages=185–207 |doi=10.1007/s00775-016-1414-3 |pmc=5350257 |pmid=27909920}}
* {{Cite journal|last1=Zurlo|first1=Giada|last2=Guo|first2=Jianping|last3=Takada|first3=Mamoru|last4=Wei|first4=Wenyi|last5=Zhang|first5=Qing|date=December 2016|title=New Insights into Protein Hydroxylation and Its Important Role in Human Diseases|journal=Biochimica et Biophysica Acta (BBA) - Reviews on Cancer|volume=1866|issue=2|pages=208–220 |doi=10.1016/j.bbcan.2016.09.004|issn=0006-3002 |pmc=5138100|pmid=27663420}}
* {{Citation |author=T. Shantha Raju |chapter=Hydroxylation of Proteins|date=2019|title=Co- and Post-Translational Modifications of Therapeutic Antibodies and Proteins|pages=119–131|publisher=John Wiley & Sons |language=en|doi=10.1002/9781119053354.ch10|isbn=978-1-119-05335-4}}
* {{cite journal |doi=10.1016/S1074-5521(99)80003-9 |pmid=10021421 |title=A hyperstable collagen mimic |journal=Chemistry & Biology |volume=6 |issue=2 |pages=63–70 |year=1999 |last1=Holmgren |first1=Steven K |last2=Bretscher |first2=Lynn E |last3=Taylor |first3=Kimberly M |last4=Raines |first4=Ronald T |doi-access=free }}
* {{cite journal |vauthors=Hausinger RP | s2cid = 85784668 | title = Fe(II)/α-ketoglutarate-dependent hydroxylases and related enzymes | journal = Crit. Rev. Biochem. Mol. Biol. | volume = 39 | issue = 1 | pages = 21–68 |date= 2004 | pmid =15121720 | doi = 10.1080/10409230490440541 }}
* {{Cite book|last1=Voet|first1=Donald|title=Principles of Biochemistry|last2=Voet|first2=Judith G.|last3=Pratt|first3=Charlotte W.|publisher=Wiley|year=2016|isbn=978-1-119-45166-2|pages=143}}
{{reflist|}}


== Спољашње везе ==
== Спољашње везе ==
{{Commonscat|Organic hydrolysis reactions}}
{{Commonscat|Organic hydrolysis reactions}}


{{Authority control}}


[[Категорија:Хемијске реакције]]
[[Категорија:Хемијске реакције]]

Верзија на датум 8. новембар 2021. у 03:07

Генерички механизам за хидролошку реакцију

Хидролиза је хемијска реакција која се заснива на распаду молекула хемијских једињења на два мања фрагмента под утицајем контакта са водом или воденом паром. The term is used broadly for substitution, elimination, and solvation reactions in which water is the nucleophile.[1] У посебном случају хидролизе соли, то је реакција јона са водом (углавном се ствара кисела или базна средина). Хидролиза је посебна грана солволизе.

Biological hydrolysis is the cleavage of biomolecules where a water molecule is consumed to effect the separation of a larger molecule into component parts. When a carbohydrate is broken into its component sugar molecules by hydrolysis (e.g., sucrose being broken down into glucose and fructose), this is recognized as saccharification.[2]

Hydrolysis reactions can be the reverse of a condensation reaction in which two molecules join into a larger one and eject a water molecule. Thus hydrolysis adds water to break down, whereas condensation builds up by removing water.[3]

Типови

Usually hydrolysis is a chemical process in which a molecule of water is added to a substance. Sometimes this addition causes both substance and water molecule to split into two parts. In such reactions, one fragment of the target molecule (or parent molecule) gains a hydrogen ion. It breaks a chemical bond in the compound.

Salts

A common kind of hydrolysis occurs when a salt of a weak acid or weak base (or both) is dissolved in water. Water spontaneously ionizes into hydroxide anions and hydronium cations. The salt also dissociates into its constituent anions and cations. For example, sodium acetate dissociates in water into sodium and acetate ions. Sodium ions react very little with the hydroxide ions whereas the acetate ions combine with hydronium ions to produce acetic acid. In this case the net result is a relative excess of hydroxide ions, yielding a basic solution.

Strong acids also undergo hydrolysis. For example, dissolving sulfuric acid (H2SO4) in water is accompanied by hydrolysis to give hydronium and bisulfate, the sulfuric acid's conjugate base. For a more technical discussion of what occurs during such a hydrolysis, see Brønsted–Lowry acid–base theory.

Esters and amides

Acid–base-catalysed hydrolyses are very common; one example is the hydrolysis of amides or esters. Their hydrolysis occurs when the nucleophile (a nucleus-seeking agent, e.g., water or hydroxyl ion) attacks the carbon of the carbonyl group of the ester or amide. In an aqueous base, hydroxyl ions are better nucleophiles than polar molecules such as water. In acids, the carbonyl group becomes protonated, and this leads to a much easier nucleophilic attack. The products for both hydrolyses are compounds with carboxylic acid groups.

Perhaps the oldest commercially practiced example of ester hydrolysis is saponification (formation of soap). It is the hydrolysis of a triglyceride (fat) with an aqueous base such as sodium hydroxide (NaOH). During the process, glycerol is formed, and the fatty acids react with the base, converting them to salts. These salts are called soaps, commonly used in households.

In addition, in living systems, most biochemical reactions (including ATP hydrolysis) take place during the catalysis of enzymes. The catalytic action of enzymes allows the hydrolysis of proteins, fats, oils, and carbohydrates. As an example, one may consider proteases (enzymes that aid digestion by causing hydrolysis of peptide bonds in proteins). They catalyze the hydrolysis of interior peptide bonds in peptide chains, as opposed to exopeptidases (another class of enzymes, that catalyze the hydrolysis of terminal peptide bonds, liberating one free amino acid at a time).

However, proteases do not catalyze the hydrolysis of all kinds of proteins. Their action is stereo-selective: Only proteins with a certain tertiary structure are targeted as some kind of orienting force is needed to place the amide group in the proper position for catalysis. The necessary contacts between an enzyme and its substrates (proteins) are created because the enzyme folds in such a way as to form a crevice into which the substrate fits; the crevice also contains the catalytic groups. Therefore, proteins that do not fit into the crevice will not undergo hydrolysis. This specificity preserves the integrity of other proteins such as hormones, and therefore the biological system continues to function normally.

Mechanism for acid-catalyzed hydrolysis of an amide.

Upon hydrolysis, an amide converts into a carboxylic acid and an amine or ammonia (which in the presence of acid are immediately converted to ammonium salts). One of the two oxygen groups on the carboxylic acid are derived from a water molecule and the amine (or ammonia) gains the hydrogen ion. The hydrolysis of peptides gives amino acids.

Many polyamide polymers such as nylon 6,6 hydrolyze in the presence of strong acids. The process leads to depolymerization. For this reason nylon products fail by fracturing when exposed to small amounts of acidic water. Polyesters are also susceptible to similar polymer degradation reactions. The problem is known as environmental stress cracking.

ATP

Hydrolysis is related to energy metabolism and storage.[4][5] All living cells require a continual supply of energy for two main purposes: the biosynthesis of micro and macromolecules, and the active transport of ions and molecules across cell membranes. The energy derived from the oxidation of nutrients is not used directly but, by means of a complex and long sequence of reactions, it is channeled into a special energy-storage molecule, adenosine triphosphate (ATP). The ATP molecule contains pyrophosphate linkages (bonds formed when two phosphate units are combined) that release energy when needed. ATP can undergo hydrolysis in two ways: Firstly, the removal of terminal phosphate to form adenosine diphosphate (ADP) and inorganic phosphate, with the reaction:

ATP + H
2
O
→ ADP + Pi

Secondly, the removal of a terminal diphosphate to yield adenosine monophosphate (AMP) and pyrophosphate. The latter usually undergoes further cleavage into its two constituent phosphates. This results in biosynthesis reactions, which usually occur in chains, that can be driven in the direction of synthesis when the phosphate bonds have undergone hydrolysis.

Полисахариди

Сахароза.[6] The glycoside bond is represented by the central oxygen atom, which holds the two monosaccharide units together.

Monosaccharides can be linked together by glycosidic bonds, which can be cleaved by hydrolysis. Two, three, several or many monosaccharides thus linked form disaccharides, trisaccharides, oligosaccharides, or polysaccharides, respectively. Enzymes that hydrolyze glycosidic bonds are called "glycoside hydrolases" or "glycosidases".

The best-known disaccharide is sucrose (table sugar). Hydrolysis of sucrose yields glucose and fructose. Invertase is a sucrase used industrially for the hydrolysis of sucrose to so-called invert sugar. Lactase is essential for digestive hydrolysis of lactose in milk; many adult humans do not produce lactase and cannot digest the lactose in milk.

The hydrolysis of polysaccharides to soluble sugars can be recognized as saccharification.[2] Malt made from barley is used as a source of β-amylase to break down starch into the disaccharide maltose, which can be used by yeast to produce beer. Other amylase enzymes may convert starch to glucose or to oligosaccharides. Cellulose is first hydrolyzed to cellobiose by cellulase and then cellobiose is further hydrolyzed to glucose by beta-glucosidase. Ruminants such as cows are able to hydrolyze cellulose into cellobiose and then glucose because of symbiotic bacteria that produce cellulases.

Metal aqua ions

Metal ions are Lewis acids, and in aqueous solution they form metal aquo complexes of the general formula M(H2O)nm+.[7][8] The aqua ions undergo hydrolysis, to a greater or lesser extent. The first hydrolysis step is given generically as

M(H2O)nm+ + H2O ⇌ M(H2O)n−1(OH)(m−1)+ + H3O+

Thus the aqua cations behave as acids in terms of Brønsted–Lowry acid–base theory.[9][10] This effect is easily explained by considering the inductive effect of the positively charged metal ion, which weakens the O-H bond of an attached water molecule, making the liberation of a proton relatively easy.

The dissociation constant, pKa, for this reaction is more or less linearly related to the charge-to-size ratio of the metal ion.[11] Ions with low charges, such as Na+ are very weak acids with almost imperceptible hydrolysis. Large divalent ions such as Ca2+, Zn2+, Sn2+ and Pb2+ have a pKa of 6 or more and would not normally be classed as acids, but small divalent ions such as Be2+ undergo extensive hydrolysis. Trivalent ions like Al3+ and Fe3+ are weak acids whose pKa is comparable to that of acetic acid. Solutions of salts such as BeCl2 or Al(NO3)3 in water are noticeably acidic; the hydrolysis can be suppressed by adding an acid such as nitric acid, making the solution more acidic.

Hydrolysis may proceed beyond the first step, often with the formation of polynuclear species via the process of olation.[11] Some "exotic" species such as Sn3(OH)42+[12] are well characterized. Hydrolysis tends to proceed as pH rises leading, in many cases, to the precipitation of a hydroxide such as Al(OH)3 or AlO(OH). These substances, major constituents of bauxite, are known as laterites and are formed by leaching from rocks of most of the ions other than aluminium and iron and subsequent hydrolysis of the remaining aluminium and iron.

Реакције хидролизе

Обично се реакција хидролизе одвија по општем шаблону:

A-B + H2O → A-H + B-OH

мада постоји могућност и за сложеније механизме те реакције, нпр.

A-B + 2 H2O → A-OH + B-OH + H2

које се одигравају у гасовитој фази при високим температурама, или у условима електролизе

Хидролиза је супротан процес од хидролитичке кондензације - тј. реакције спајања два или више молекула са издвајањем молекула воде.[13]

Многе реакције хидролизе су повратне, при чему смер реакције зависи од њених услова:

Нпр. хидролиза естара:

RCOOR' + H2O → RCOOH + R'OH

захтева додавање одређене количине воде у реакциони систем; уколико нема довољно воде настаје реакција кондензације тек настале киселине (RCOOH) и алкохола (OH), која је у ствари иста као и претходна само супротног смера[13]:

RCOOH + R'OH → RCOOR' + H2O

Реакција хидролизе није исто што и Електролитичка дисоцијација.

Дисоцијација се заснива на распаду молекула под дејством растварача (растварач може бити и вода) али без грађења ковалентних веза са њим, док се хидратација заснива на стварању комплексних једињења са водом везаних водоничном везом.

Види још

Референце

  1. ^ IUPAC. „Hydrolysis”. Kompendijum hemijske terminologije (Internet izdanje).IUPAC. „Solvolysis”. Kompendijum hemijske terminologije (Internet izdanje).
  2. ^ а б „Definition of Saccharification”. www.merriam-webster.com (на језику: енглески). Архивирано из оригинала 7. 1. 2021. г. Приступљено 8. 9. 2020. 
  3. ^ Steane, Richard. „Condensation and Hydrolysis”. www.biotopics.co.uk. Архивирано из оригинала 2020-11-27. г. Приступљено 2020-11-13. 
  4. ^ Nelson, David L., Cox, Michael M. Lehninger: Principles of Biochemistry. New York: W.H. Freeman and Company, 2013. Sixth ed., p. 522–523.
  5. ^ Hardie, D. G., Ross, F. A., Hawley, S. A (2012). AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nature, 13, 251–262. Accessed 9 April 2017.
  6. ^ Ryan KJ, Ray CG, ур. (2004). Sherris Medical Microbiology (4th изд.). McGraw Hill. ISBN 978-0-8385-8529-0. 
  7. ^ Burgess, John (1978). Metal Ions in Solution. Chichester: Ellis Horwood. ISBN 978-0853120278. 
  8. ^ Richens, D. T. (1997). The Chemistry of Aqua Ions: Synthesis, Structure, and Reactivity: A Tour through the Periodic Table of the Elements. Wiley. ISBN 0-471-97058-1. 
  9. ^ Brönsted, J. N. (1923). „Einige Bemerkungen über den Begriff der Säuren und Basen” [Some observations about the concept of acids and bases]. Recueil des Travaux Chimiques des Pays-Bas. 42 (8): 718—728. doi:10.1002/recl.19230420815. 
  10. ^ Lowry, T. M. (1923). „The uniqueness of hydrogen”. Journal of the Society of Chemical Industry. 42 (3): 43—47. doi:10.1002/jctb.5000420302. 
  11. ^ а б Baes, Charles F.; Mesmer, Robert E. (1976). The Hydrolysis of Cations. New York: Wiley. ISBN 9780471039853. 
  12. ^ Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (II изд.). Oxford: Butterworth-Heinemann. стр. 384. ISBN 0080379419. 
  13. ^ а б McMurry John E. (1992). Fundamentals of Organic Chemistry (3rd изд.). Belmont: Wadsworth. ISBN 0-534-16218-5. 

Литература

Спољашње везе