PHSE 621 PAC

PHSE 621 PAC PHSE 621 PAC

v.sgsync.lnx1.nwu.ac.za
from v.sgsync.lnx1.nwu.ac.za More from this publisher
02.05.2013 Views

GEVORDERDE CHEMIE VIR ONDERWYS II STUDIEGIDS VIR PHSE 621 PAC *PHSE621PAC* FAKULTEIT NATUURWETENSKAPPE

GEVORDERDE CHEMIE VIR ONDERWYS II<br />

STUDIEGIDS VIR<br />

<strong>PHSE</strong> <strong>621</strong> <strong>PAC</strong><br />

*<strong>PHSE</strong><strong>621</strong><strong>PAC</strong>*<br />

FAKULTEIT NATUURWETENSKAPPE


Studiegids saamgestel deur:<br />

Dr J Röscher<br />

Bladuitleg deur Justus Röscher, Skool vir Fisiese- en Chemiese Wetenskappe<br />

Hantering van drukwerk en verspreiding deur Departement Logistiek (Verspreidingsentrum)<br />

Gedruk deur Platinum Press (018) 2994226<br />

Kopiereg 2012 uitgawe. Hersieningsdatum 2013<br />

Noordwes-Universiteit, Potchefstroomkampus<br />

Kopiereg voorbehou. Geen gedeelte van hierdie boek mag in enige vorm of op enige manier<br />

sonder skriftelike toestemming van die uitgewer weergegee word nie. Dit sluit in<br />

fotokopiëring van die boek of gedeeltes van die boek.<br />

ii


Inhoud<br />

Inligting oor die module v<br />

1. Inleiding v<br />

2. Module-uitkomste v<br />

3. Inligting oor die dosent vi<br />

4. Studiemateriaal vi<br />

4.1. Voorgeskrewe handboek vi<br />

4.2. Study material included in the study guide vi<br />

5. Aksiewoorde en die aard van nagraadse studie vii<br />

6. Werkswyse viii<br />

Afwesigheid viii<br />

7. Assesseringsinligting viii<br />

7.1. Formatiewe assessering [Kontinue assessering] viii<br />

Deelnamepunt ix<br />

Deelnamebewys ix<br />

7.2. Summatiewe assessering [Eksamen] ix<br />

Modulepunt [finale punt] ix<br />

8. Hoe om hierdie studie aan te pak ix<br />

9. Kurrikulum x<br />

10. Waarskuwing teen plagiaat xi<br />

1. Die omgewing en omgewingsprobleme 1<br />

1.1. Die omgewing 1<br />

1.1.1. Inleiding 1<br />

1.1.2. Studiemateriaal 1<br />

1.1.3. Self-evaluering 1<br />

1.2. Omgewingsprobleme, hul oorsake en volhoudbaarheid 1<br />

1.2.1. Inleiding 1<br />

1.2.2. Study material 2<br />

1.2.3. Self-evaluation 2<br />

1.3. Ekosisteme – wat dit is en hoe dit werk 2<br />

1.3.1. Studiemateriaal 2<br />

1.3.2. Self-evaluering 2<br />

2. Chemie en die omgewing 3<br />

2.1. Waarom chemie omgewingsprobleme veroorsaak 3<br />

2.2. Omgewingschemie 4<br />

2.2.1. Inleiding 4<br />

2.2.2. Studiemateriaal 4<br />

2.2.3. Self-evaluering 4<br />

2.3. Elementêre biochemie en toksikologie 4<br />

2.3.1. Studiemateriaal 5<br />

2.3.2. Self-evaluering 5<br />

2.4. Groen chemie 5<br />

2.4.1. Inleiding 5<br />

2.4.2. Studiemateriaal 5<br />

2.4.3. Self-evaluering 5<br />

3. Chemie en die atmosfeer 7<br />

3.1. Inleiding 7<br />

3.2. Die atmosfeer en lugbesoedeling 7<br />

3.2.1. Studiemateriaal 7<br />

3.2.2. Self-evaluering 8<br />

3.3. Klimaatontwrigting en osoonvernietiging 8<br />

3.3.1. Studiemateriaal 8<br />

3.3.2. Self-evaluering 8<br />

iii


4. Chemie en die hidrosfeer 9<br />

4.1. Inleiding 9<br />

4.2. Waterbesoedeling 9<br />

4.2.1. Studiemateriaal 9<br />

4.2.2. Self-evaluering 9<br />

4.3. Watersuiwering 10<br />

4.3.1. Studiemateriaal 10<br />

4.3.2. Self-evaluering 10<br />

5. Chemie en die litosfeer 11<br />

5.1. Inleiding 11<br />

5.2. Geologie en mineraalbronne 11<br />

5.2.1. Studiemateriaal 11<br />

5.2.2. Self-evaluering 12<br />

5.3. Basiese beginsels van mynbou en die ekstraksie van metale 12<br />

5.3.1. Studiemateriaal 13<br />

5.3.2. Self-evaluering 13<br />

iv


1. Inleiding<br />

Inligting oor die module<br />

Hierdie module is 'n inleiding tot omgewingschemie. Om die belang van hierdie veld te verstaan,<br />

behoort studente 'n basiese begrip van en waardering vir die omgewing te hê. Omgewingschemie is 'n<br />

multidissiplinêre veld wat gebruik maak van agtergrond vanaf ander dissiplines, soos byvoorbeeld<br />

chemie, biochemie, mikrobiologie, toksikologie, geologie, geografie, plantkunde en dierkunde. Waar<br />

moontlik sal Suid-Afrikaanse omgewingsprobleme uitgelig word, insluitende 'n bespreking van die<br />

mynboubedryf in Suid-Afrika. Die module sluit af met 'n toekomsperspektief oor die rol van chemie in<br />

die industrie en gemeenskap. Omgewingschemie verskaf 'n ryk verskeidenheid kontekste en<br />

toepassings van chemie as gevolg van die multidissiplinêre aard van die vakgebied. Kennis en begrip<br />

van omgewingschemie kan grootliks bydra tot die vermoë van die fisiese wetenskappe onderwyser om<br />

die toepassings van teoretiese chemiese beginsels in die "werklikheid" te verstaan.<br />

Omgewingschemie is multidissiplinêr en daar sal dus van die onderwyser verwag word om<br />

verskillende chemiekonsepte met mekaar te skakel en sodoende kognitiewe strukture te bou wat<br />

nodig is vir helder begrip. Die kennis en vaardighede wat in hierdie module verwerf kan word, sal die<br />

fisiese wetenskappe onderwyser in staat stel om chemie teorie in 'n omgewingskonteks te plaas en<br />

sodoende die leerders help om die relevansie en toepasbaarheid van chemie te begryp.<br />

Hierdie module aanvaar dat u 'n basiese agtergrond van chemie in u vorige studiejare verwerf het. Die<br />

kredietwaarde van die module is 16 en daar word dus verwag dat u ongeveer 160 ure benodig om die<br />

studie te voltooi. Hierdie geskatte studietyd sluit voorbereiding, praktiese werk en assesseringstake in.<br />

Ek hoop dat hierdie module u begrip van chemie in konteks merkbaar sal vergroot.<br />

J. Röscher<br />

April 2012<br />

2. Module-uitkomste<br />

Na voltooiing van hierdie module behoort studente die volgende te kan aantoon.<br />

Omvattende kennis en insig oor 'n wye verskeidenheid nuwe asook bekende konsepte, beginsels,<br />

wette, teorieë en modelle wat deel uitmaak van chemie in die skoolkurrikulum of wat die<br />

effektiewe ontwikkeling van hierdie vakgebied ondersteun.<br />

Die vermoë om probleme binne die multidissiplinêre veld van omgewingschemie op te los deur<br />

kennis van uiteenlopende temas binne chemie en ander vakgebiede te integreer.<br />

Die vermoë om leerders se idees oor geselekteerde konsepte wat omgewingschemie onderlê te<br />

analiseer en evalueer en hierdie idees te verfyn deur implementering van konstruktivistiese<br />

onderrig.<br />

Die vermoë om by te dra tot kritiese en sistematiese denke oor die invloed van wetenskap en<br />

tegnologie op die samelewing en die omgewing.<br />

Basiese laboratoriumvaardigheid en die vermoë om effektiewe praktiese aktiwiteite op skoolvlak te<br />

beplan en aan te bied.<br />

v


Inligting oor die module<br />

3. Inligting oor die dosent<br />

vi<br />

Naam van dosent J Röscher<br />

Kantoornommer en gebou Kantoor G16, Gebou G5<br />

Telefoonnommer 018 299 2333<br />

Faksnommer 086 686 4997<br />

E-pos 10091858@nwu.ac.za<br />

4. Studiemateriaal<br />

4.1. Voorgeskrewe handboek<br />

Miller, GT and Spoolman, SE, 2012, Living in the environmental, 17 th ed., Brooks/Cole.<br />

4.2. Studiemateriaal ingesluit by hierdie studiegids<br />

Uittreksels van die volgende boeke is ingesluit in die bylae aan die einde van die studiegids.<br />

Bylaag 1: Manahan, SE, 2003, Toxicological chemistry and biochemistry, CRC Press, Boca<br />

Raton.<br />

Bylaag 2: Clark, JH and Macquarrie, D, 2002, Handbook of Green Chemistry and Technology,<br />

2 nd ed., Blackwell Science: Oxford.<br />

Bylaag 3: Baird, C and Cann, M, 2008, Environmental chemistry, 4 th ed., W. H. Freeman and<br />

Company: New York, NY.<br />

Bylaag 4: Matthews, P, 1992, Advanced chemistry, Cambridge: Chennaice: Oxford.<br />

Bylaag 5: Hill, JW, Petrucci, RH, McCreary, TW and Perry, SS, 2005, General chemistry, 4 th ed.,<br />

Pearson: Upper Saddle River, NY.


5. Aksiewoorde en die aard van nagraadse studie<br />

Inligting oor die module<br />

Volgens Bloom kan kognitiewe vaardighede in 'n hiërargiese stelsel georganiseer word wat strek<br />

vanaf kennis aan die onderpunt tot evaluering aan die bopunt via begrip, toepassing, ontleding en<br />

sintese. In nagraadse studie verskuif die klem van studie na hoërorde kognitiewe vaardighede.<br />

Gevolglik vereis studie-aktiwiteite die geïntegreerde gebruik van meerdere vlakke van kognitiewe<br />

vaardighede deur die student. Die aksiewerkwoorde wat in hierdie module gebruik word, is ‘n<br />

aanduiding van die kognitiewe vlak van die betrokke handeling, soos onder aangedui:<br />

Kognitiewe vlak Aksiewoorde Verduideliking<br />

VLAK 1:<br />

KENNIS<br />

VLAK 2:<br />

INSIG<br />

VLAK 3:<br />

TOEPASSING<br />

Definieer<br />

Beskryf<br />

Lys, Skryf, Gee, Noem<br />

Verduidelik, Bespreek<br />

Illustreer<br />

Onderskei, Vergelyk<br />

Som op<br />

Bepaal<br />

Bereken<br />

Stel meganisme voor<br />

VLAK 4:<br />

ANALISE Analiseer, Bespreek<br />

VLAK 5:<br />

SINTESE<br />

VLAK 6:<br />

EVALUERING<br />

Bereken<br />

Bewys<br />

Dui verband aan<br />

Som op of konstrueer<br />

Kritiseer, Evaluering<br />

Gee 'n akkurate, kort beskrywing van 'n begrip sodat die<br />

betekenis daarvan duidelik blyk.<br />

Eienskappe, feite of resultate word op ’n logiese, goed<br />

geformuleerde wyse weergegee. Geen bespreking of<br />

verduideliking is nodig nie.<br />

Gee die antwoord (feite) puntsgewys. Geen bespreking<br />

of verduideliking is nodig nie.<br />

Gee redes op ’n logiese, goed gestruktureerde wyse<br />

vanuit illustrasies, modelle, wette en wiskundige<br />

vergelykings.<br />

Beskryf 'n begrip met behulp van voorbeelde of 'n skets<br />

of diagram met of sonder byskrifte.<br />

Feite, gebeure of probleme word teenoor mekaar gestel<br />

en ooreenkomste en verskille word na vore gebring.<br />

Om die wesenlike inligting op 'n beknopte en<br />

sistematiese manier weer te gee.<br />

Pas bestaande kennis en metodes (strategieë) toe op 'n<br />

nuwe probleem of situasie.<br />

Enkele wiskundige metodes word toegepas om 'n<br />

numeriese antwoord te kry.<br />

Gee ’n meganisme, d.w.s. reaksieverloop, met<br />

pyltjienotasie en tussenstappe.<br />

Verdeel 'n probleem, stelling of idee in die<br />

samestellende dele. Verduidelik die belang van elke<br />

deel en dui die onderlinge verwantskap tussen dele<br />

aan.<br />

Meerdere wiskundige metodes word toegepas om 'n<br />

numeriese antwoord te kry.<br />

Stellings word deur logiese aanvoer van aanvaarbare<br />

feite gestaaf.<br />

Vind en verduidelik die verwantskap tussen verskillende<br />

stellings.<br />

'n Groot massa kennis word opgesom en logies en<br />

sistematies georganiseer terwyl die essensie van die<br />

saak behou word.<br />

Bepaal die waarde van 'n stelling, kwessie of argument<br />

deur te verduidelik of jy daarmee saamstem of verskil.<br />

Gee redes vir jou opinies. Analiseer die probleem en<br />

bepaal die waarde van elke komponent. Die resultaat<br />

word opgesom (sintese) om 'n omvattende en<br />

doelgerigte waardebeoordeling te lewer.<br />

vii


Inligting oor die module<br />

6. Werkswyse<br />

Omvattende selfstudie word van nagraadse studente verwag. Die kontaktyd sal primêr gebruik word<br />

om inleidende raamwerke vir die module-inhoude te konstrueer wat later as deel van verdere studie<br />

gebruik kan gebruik. Tyd sal ook afgestaan word aan klasbesprekings en om oefeninge te voltooi. Die<br />

tema van hierdie module is breed en dit is nie moontlik om al die inhoude in die klas te "behandel" nie.<br />

Daarom is dit belangrik dat u voorbereid klas toe kom en selfstudie doen as dit van u verwag word.<br />

Oopboekbenadering<br />

Praktiese werk<br />

Velduitstappie<br />

viii<br />

Die dosent sal u inlig of 'n oopboekbenadering in hierdie module gevolg<br />

word. Lees asseblief die afdeling "Hoe om hierdie studie aan te pak" vir<br />

meer inligting oor hierdie benadering.<br />

Verpligte praktiese werk vorm deel van hierdie module. Die rooster vir<br />

praktiese werk sal aan die begin van die semester beskikbaar gestel word.<br />

'n Verpligte velduitstappie kan moontlik as deel van die module<br />

georganiseer word. Inligting sal vroegtydig aan studente deurgegee word.<br />

Kontak tussen studente en tussen studente en die dosent dra beduidend by tot die betekenisvolle leer<br />

van chemie en fisika. U word aangemoedig en daar word van u verwag om al die kontaksessies by te<br />

woon. Raadpleeg asseblief die prospektus vir verdere inligting oor die kontakgeleenthede in 2009.<br />

Afwesigheid<br />

Indien u nie 'n toets kon skryf nie, moet u so gou as moontlik 'n mediese sertifikaat of ander amptelike<br />

stawende dokumentasie aan die dosent verskaf. Geen verskoning word aangeteken indien die<br />

nodige dokumentasie nie voorsien is nie. In die geval van afwesigheid moet u die moontlikheid van<br />

'n alternatiewe assesseringsgeleentheid of ingeedatum met die dosent bespreek. Geen alternatiewe<br />

rëelings kan in die geval van praktiese werk of velduitstappies getref word nie.<br />

7. Assesseringsinligting<br />

7.1. Formatiewe assessering [Kontinue assessering]<br />

Assesseringsgeleenthede wat bydra tot die deelnamepunt word in die tabel gelys.<br />

Assessering Beskrywing<br />

Klastoetse<br />

Praktiese verslae<br />

Opdragte<br />

Klastoetse word geskryf om u kennis en insig aangaande<br />

die huidige inhoud en die onderliggende chemie teorie<br />

(wat u in vorige chemiemodules moes bemeester het) te<br />

toets. Die datums vir klastoetse sal in die klas bekend<br />

gemaak word.<br />

Praktiese verslaggewing kan aspekte soos die invul van<br />

verslagblaaie, saamstel van 'n omvattende verslag (soos<br />

byvoorbeeld oor 'n velduitstappie) of die ontwikkeling van<br />

'n praktiese aktiwiteit vir gebruik op 'n sekere skoolvlak<br />

behels. Ingeedatums sal in die klas bekend gemaak word.<br />

Opdragte wat daarop gemik is om u te help om spesifieke<br />

vaardighede te bemeester, sal somtyds gegee word.<br />

Ingeedatums sal in die klas bekend gemaak word.<br />

Bydrae tot<br />

deelnamepunt<br />

40%<br />

30%<br />

30%


Deelnamepunt<br />

Die deelnamepunt word as volg saamgestel:<br />

Deelnamepunt = toetse (40%) + praktiese verlae (30%) + opdragte (30%)<br />

Deelnamebewys<br />

Inligting oor die module<br />

Om toelating tot die eksamen te verkry, moet u 'n deelnamebewys verwerf. 'n Deelnamebewys sal<br />

toegeken word indien u deelnamepunt van minstens 40% behaal het.<br />

7.2. Summatiewe assessering [Eksamen]<br />

'n Vraestel van 100 punte en 3 ure word aan die einde van die semester afgelê. 'n Subminimum van<br />

40% moet in die eksamen behaal word ongeag die finale punt wat in die module behaal word.<br />

Modulepunt [finale punt]<br />

Die modulepunt word as volg saamgestel:<br />

Modulepunt = deelnamepunt ( 1 /2) + eksamenpunt ( 1 /2).<br />

'n Modulepunt van ten minste 50% moet verwerf word om die module te slaag.<br />

8. Hoe om hierdie studie aan te pak<br />

Die volgende drie basiese riglyne kan jou help om hierdie module suksesvolle te voltooi.<br />

1. Basiese kennis. Hierdie module bou op die fondament wat tydens u voorgraadse<br />

chemiemodules gelê is. U moet dus reeds 'n basiese begrip van die belangrikste temas in die<br />

vakgebied ontwikkel het.<br />

2. Effektiewe tydsbestuur. Maak seker dat u genoeg geleenthede het om te studeer en dat hierdie<br />

geleenthede goed versprei en lank genoeg is. Tydsbestuur is veral belangrik tydens deeltydse<br />

studie.<br />

ix


Inligting oor die module<br />

3. Korrekte studiemetodes. Die hoeveelheid studiemateriaal in hierdie module is aansienlik.<br />

Omdat dit nie moontlik is om alles te leer nie, is effektiewe studiestrategieë noodsaaklik.<br />

Dit is belangrik om die belangrikste dele van elke afdeling te kan identifiseer sodat die meeste<br />

studietyd daaraan gewy kan word. In hierdie module is toepassing, interpretasie en die<br />

identifikasie van die verwantskappe tussen temas belangriker as blote feite.<br />

Goeie, sistematiese opsommings is baie handig aangesien dit help om groot volumes werk te<br />

hanteer. Gebruik tabelle en diagramme om die logiese samehang tussen verskillende dele van die<br />

studiemateriaal aan te toon. Logiese studiemateriaal leer makliker!<br />

Net soos op skool is oefening en herhaling die beste metodes om te leer. Daar is talle probleme<br />

met antwoorde aan die einde van elke leereenheid in die voorgeskrewe handboek. U moet<br />

besonder versigtig wees as 'n oopboekbenadering in hierdie module gevolg word. Die effektiewe<br />

gebruik van hierdie benadering is 'n vaardigheid wat aangeleer en geslyp moet word deur<br />

voldoende en gereelde oefening. 'n Oopboekbenadering beteken definitief nie dat u nie hoef te<br />

leer nie.<br />

4. Stiptelikheid en toewyding aan studies. Woon alle kontakgeleenthede by en skryf al die toetse.<br />

Voltooi en handig oefeninge, verslae en opdragte betyds in. Statistiek toon duidelik dat<br />

toegewyde studente 'n verbeterde kans het om hul studies suksesvol te voltooi.<br />

9. Kurrikulum<br />

Kurrikulumstruktuur vir Fisiese Wetenskappe Onderwys (Deeltyds)<br />

JAAR 1<br />

Semester 1 Semester 2<br />

x<br />

Module<br />

kode<br />

FOER 611<br />

TLAS 612<br />

<strong>PHSE</strong> 612<br />

Module name Krediete<br />

Grondslae van<br />

onderrignavorsing<br />

Onderrig, leer en<br />

assessering<br />

Gevorderde chemie<br />

vir onderwys I<br />

Module<br />

kode<br />

16 <strong>PHSE</strong> 622<br />

16<br />

8<br />

Module name Krediete<br />

Gevorderde fisika<br />

vir onderwys II<br />

Totaal 40 Totaal 16<br />

JAAR 2<br />

Semester 3 Semester 4<br />

Module<br />

kode<br />

Module name Krediete<br />

Module<br />

kode<br />

16<br />

Module name Krediete<br />

RSPR 671* Navorsingsprojek 16 RSPR 671* Navorsingsprojek 16<br />

CUDE 611 Kurrikumontwikkeling 16 <strong>PHSE</strong> <strong>621</strong><br />

<strong>PHSE</strong> 611<br />

Gevorderde fisika<br />

vir onderwys I<br />

8<br />

Gevorderde chemie<br />

vir onderwys II<br />

Totaal 40 Totaal 32<br />

* 32-krediet module aangebied oor drie semesters<br />

TOTAAL 128<br />

16


Kurrikulumstruktuur vir Fisiese Wetenskappe Onderwys (Voltyds)<br />

JAAR 1<br />

Semester 1 Semester 2<br />

Module<br />

kode<br />

FOER 611<br />

TLAS 612<br />

Module name Krediete<br />

Grondslae van<br />

onderrignavorsing<br />

Onderrig, leer en<br />

assessering<br />

Module<br />

kode<br />

Inligting oor die module<br />

Module name Krediete<br />

16 RSPR 671 Navorsingsprojek 32<br />

16 <strong>PHSE</strong> <strong>621</strong><br />

CUDE 611 Kurrikumontwikkeling 16 <strong>PHSE</strong> 622<br />

<strong>PHSE</strong> 611<br />

<strong>PHSE</strong> 612<br />

Gevorderde fisika<br />

vir onderwys I<br />

Gevorderde chemie<br />

vir onderwys I<br />

8<br />

8<br />

Gevorderde chemie<br />

vir onderwys II<br />

Gevorderde fisika<br />

vir onderwys II<br />

Totaal 72 Totaal 56<br />

TOTAAL 128<br />

10. Waarskuwing teen plagiaat<br />

Werkstukke is individuele take en nie groepaktiwiteite nie (tensy dit uitdruklik aangedui word<br />

as 'n groepaktiwiteit)<br />

Kopiëring van teks van ander leerders of uit ander bronne (byvoorbeeld die studiegids, voorgeskrewe<br />

studiemateriaal of direk vanaf die internet) is ontoelaatbaar – net kort aanhalings is toelaatbaar en<br />

slegs indien dit as sodanig aangedui word.<br />

U moet bestaande teks herformuleer en u eie woorde gebruik om te verduidelik wat u gelees het. Dit<br />

is nie aanvaarbaar om bestaande teks/stof/inligting bloot oor te tik en die bron in 'n voetnoot te erken<br />

nie – u behoort in staat te wees om die idee of begrip/konsep weer te gee sonder om die<br />

oorspronklike skrywer woordeliks te herhaal.<br />

Die doel van die opdragte is nie die blote weergee van bestaande materiaal/stof nie, maar om vas te<br />

stel of u oor die vermoë beskik om bestaande tekste te integreer, om u eie interpretasie en/of kritiese<br />

beoordeling te formuleer en om 'n kreatiewe oplossing vir bestaande probleme te bied.<br />

WEES GEWAARSKU: Studente wat gekopieerde teks indien sal 'n nulpunt vir die opdrag<br />

ontvang en dissiplinêre stappe mag deur die Fakulteit en/of die Universiteit teen sodanige<br />

studente geneem word. Dit is ook onaanvaarbaar om iemand anders se werk vir hulle te doen<br />

of iemand anders in staat te stel om u werk te kopieer – moet dus nie u werk uitleen of<br />

beskikbaar stel aan ander nie!<br />

16<br />

16<br />

xi


1. Die omgewing en omgewingsprobleme<br />

TYD BENODIG Dit sal ongeveer 32 ure neem om hierdie leereenheid te voltooi.<br />

UITKOMSTE Gee 'n oorsig van die oorsprong en ontwikkeling van die aarde.<br />

Definieer die term "volhoubaarheid".<br />

Verduidelik die beginsels van volhoubaarheid.<br />

Bespreek die begrip "ekologiese voetspoor".<br />

Verduidelik die "tragedy of the commons" en illustreer die idee met<br />

voorbeelde.<br />

Beskryf en bespreek die IPAT omgewingsinpakmodel.<br />

Bespreek die redes vir ons omgewingsprobleme.<br />

1.1. Die omgewing<br />

1.1.1. Inleiding<br />

"There is a sufficiency in the world for man’s need but not for man’s greed."<br />

Mahatma Gandhi (1869–1948)<br />

Ons lewe in 'n besondere komplekse omgewing wat in staat is om lewe te onderhou. In hierdie<br />

afdeling word studente met min of geen agtergrond van die omgewingswetenskappe begelei tot 'n<br />

basiese begrip van die ontwikkeling van die aarde en 'n waardering vir die interafhanklikheid van alle<br />

elemente van die omgewing.<br />

1.1.2. Studiemateriaal<br />

(a) DVD: How the earth was made.<br />

(b) Verdere notas mag moontlik in die klas verskaf word.<br />

1.1.3. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

1.2. Omgewingsprobleme, hul oorsake en volhoubaarheid<br />

1.2.1. Inleiding<br />

Omgewingswetenskap is 'n multidissiplinêre vakgebied wat handel oor hoe die aarde werk, hoe ons<br />

met die aarde omgaan en hoe ons omgewingsprobleme kan hanteer. Omgewingsvraagstukke<br />

1


Die omgewing en omgewingsprobleme<br />

beïnvloed elke deel van ons lewens en daarom is die inhoud van die voorgeskrewe handboek en<br />

hierdie kursus uiters relevant vir elke student. Die skrywers van die handboek beskou<br />

omgewingswetenskap as die belangrikste kursus in 'n student se opvoeding. Hulle voer aan dat min<br />

sake so belangrik kan wees as om te leer hoe die aarde werk en van die mens se invloed op die<br />

aarde se vermoë om ons te onderhou. Omgewingswetenskap bied ook voorstelle aan oor hoe ons<br />

omgewingsinpak kan verminder. Ons lewe in 'n uitdagende era en dit is duidelik dat ons in hierdie eeu<br />

ingrypende kultuurveranderings sal moet maak om meer volhoubaar te lewe.<br />

1.2.2. Study material<br />

(a) Environmental problems, their causes, and sustainability, [Living in the environment], p. 5 – 30.<br />

1.2.3. Self-evaluation<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

1.3. Ekosisteme – wat dit is en hoe dit werk<br />

1.3.1. Studiemateriaal<br />

(a) Ecosystems: what are they and how do they work? [Living in the environment], p. 54 – 79.<br />

1.3.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

2


2. Chemie en die omgewing<br />

TYD BENODIG Dit sal ongeveer 32 ure neem om hierdie leereenheid te voltooi.<br />

UITKOMSTE Verduidelik waarom die manier waarop die mens chemie aanwend (en<br />

misbruik) uiteindelik tot omgewingsprobleme aanleiding sal gee.<br />

Definieer die term "omgewingchemie".<br />

Gee 'n oorsig van basiese biochemie.<br />

Beskryf van die kategorieë waarin toksisiteit verdeel kan word.<br />

Vergelyk ou en nuwe maniere om oor chemie te dink.<br />

Gee redes waarom groen chemie belangrik is vir die samelewing en industrie.<br />

Lys die beginsels van groen chemie.<br />

Verduidelik die beginsel van atoomekonomie en bespreek die bruikbaarheid<br />

daarvan.<br />

Berekening van die atoomekonomie van gegewe reaksies.<br />

Gee voorbeelde van industriële vervaardigingsmetodes wat meer "groen"<br />

gemaak is.<br />

2.1. Waarom chemie omgewingsprobleme veroorsaak<br />

'n Fundamentele oorsaak van ons groterwordende omgewingsprobleme is dat mense chemie<br />

anders bedryf as bestaande natuurlike prosesse.<br />

For eons, biochemical processes have evolved by drawing primarily on elements that are abundant<br />

and close at hand – such as carbon, hydrogen, oxygen, nitrogen, sulphur, calcium and iron – to create<br />

everything from paramecia to redwoods, clown fish to humans. Our industries, in contrast, gather<br />

elements from nearly every corner of the planet and distribute them in ways natural processes never<br />

could. Lead, for example, used to be found mostly in deposits so isolated and remote that nature<br />

never folded it into living organisms. But now lead is everywhere, primarily because our paints, cars<br />

and computers have spread it around. When it finds its way into children, even at minuscule doses, it<br />

is severely toxic. The same can be said for arsenic, cadmium, mercury, uranium and plutonium. These<br />

elements are persistent pollutants – they do not degrade in animal bodies or in the surrounding<br />

environment – so there is a pressing need to find safer alternatives. Some of the new synthetic<br />

molecules in medicines, plastics and pesticides are so different from the products of natural chemistry<br />

that it is as though they dropped in from an alien world. Many of these molecules do not degrade<br />

easily, and even some biodegradable compounds have become omnipresent because we use them<br />

so copiously.<br />

Recent research indicates that some of these substances can interfere with the normal expression of<br />

genes involved in the development of the male reproductive system. Scientists have known for several<br />

years that prenatal exposure to phthalates (compounds used in plastics and beauty products) can<br />

alter the reproductive tract of new-born male rodents and in 2005 similar effects were reported in male<br />

infants. Another study found that men with low sperm counts living in a rural farming area of Missouri<br />

had elevated levels of herbicides (such as alachlor and atrazine) in their urine. Starting from our<br />

factories, farms and sewers, persistent pollutants can journey intact by air, water and up the food<br />

chain, often right back to us.<br />

[Aangepas vanuit: Environmental chemistry]<br />

3


Chemie en die omgewing<br />

2.2. Omgewingschemie<br />

2.2.1. Inleiding<br />

Chemie is die studie van materie. Daarom het dit te make met die lug wat ons inasem, die water wat<br />

ons drink en die grond waarin ons graan verbou. Mense, plante en diere bevat 'n ongelooflike<br />

hoeveelheid chemiese stowwe wat verwant is deur 'n groot aantal komplekse chemiese reaksies.<br />

Daar is vandag groot besorgdheid oor die gebruik en misbruik van chemie en die invloed wat hierdie<br />

aktiwiteite op die omgewing het. Verskeie gebeure in die geskiedenis van die mensdom herinner ons<br />

aan die gevaar wat die mens vir die omgewing inhou – van individuele blootstelling aan skadelike<br />

stowwe tot verskynsels op 'n globale skaal wat moontlike katastrofiese gevolge kan hê.<br />

Gedurende die 20 ste eeu was die publiek nie werklik bewus van die omgewingasprobleme wat besig<br />

was om te onstaan nie. Die sentiment van die publiek het egter verander aan die begin van die 21 ste<br />

eeu deurdat wetenskaplikes en omgewingsorganisasies die boodskap begin versprei het dat<br />

omgewingsprobleme die volhoubaarheid van die aarde in gedrang bring. Vandag is weinig mense nie<br />

bewus van die gevare wat die kwaliteit van ons lug, water en grond bedreig nie. Kennis bring die<br />

verantwoordelikheid om oplossings te vind vir die probleme wat deur die groeiende menslike bevolking<br />

veroorsaak word. Om omgewingsprobleme te verstaan en oplossings te vind vereis kennis van<br />

omgewingschemie en natuurlik ook ander omgewingswetenskappe. Omgewingschemie is die<br />

afdeling van chemie wat die oorsprong, transport, reaksies, effek en uiteinde van chemiese stowwe in<br />

die lug, water grond en lewende sisteme bestudeer, insluitende die invloed van menslike aktiwiteite<br />

hierop. 'n Verwante dissipline is toksikologiese chemie wat die chemie van gifstowwe bestudeer met<br />

die klem op die interaksie van gifstowwe met biologiese weefsel en lewende sisteme.<br />

2.2.2. Studiemateriaal<br />

(a) [Toxicological chemistry and biochemistry], p. 39 – 75. Sien Bylaag 1.<br />

2.2.3. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

2.3. Elementêre biochemie en toksikologie<br />

GEVALLESTUDIE: DDT in moedersmelk in KZN<br />

Fokusvrae<br />

1. Wat is die oorsprong van die DDT in die moedersmelk?<br />

2. Waarom verbied ons nie bloot die gebruik van DDT nie? Waarom is die verbanning van DDT 'n<br />

emosionele saak?<br />

3. Waarom kom die DDT spesifiek in moedersmelk voor? In watter ander deel of dele van die<br />

liggaam kan DDT ook verwag word?<br />

4. Wat is bioamplifikasie?<br />

5. Waarom is DDT giftig vir mense?<br />

4


2.3.1. Studiemateriaal<br />

(a) [Toxicological chemistry and biochemistry], p. 39 – 75. See Bylaag 1.<br />

(b) Environmental hazards and human health, [Living in the environment], p. 436 – 464.<br />

(c) Verdere notas mag moontlik in die klas verskaf word.<br />

2.3.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

2.4. Groen chemie<br />

2.4.1. Inleiding<br />

Chemie en die omgewing<br />

Chemie het dit tans maar moeilik. Die samelewing verlang groter hoeveelhede van toenemend meer<br />

komplekse chemiese stowwe, maar bejeën ook die industrieë wat hierdie produkte lewer met meer en<br />

meer vrees en agterdog. Die idees van groen chemie kan moontlik oplossings verskaf wat alle partye<br />

tevrede stel. Hierdie afdeling verduidelik die beginsels van groen chemie en gee voorbeelde van die<br />

toepassing daarvan.<br />

2.4.2. Studiemateriaal<br />

(a) [Handbook of Green Chemistry and Technology], p. 1 – 28. See Bylaag 2.<br />

2.4.3. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

5


3. Chemie en die atmosfeer<br />

TYD BENODIG Dit sal ongeveer 32 ure neem om hierdie leereenheid te voltooi.<br />

UITKOMSTE Beskryf die samestelling van die atmosfeer.<br />

Bespreek die belangrikste buitenshuise lugbesoedelingsprobleme.<br />

3.1. Inleiding<br />

Bespreek van die belangrikste binnenshuise lugbesoedelingsprobleme.<br />

Bespreek die effek van lugbesoedeling op gesondheid.<br />

Verduidelik die funksie van osoon in die atmosfeer, die meganisme van<br />

osoonvernietiging en stappe wat geneem is om die probleem op te los.<br />

Bespreek die aardverwarmingsprobleem.<br />

Bespreek die suurreënprobleem.<br />

Stel wyses voor om lugbesoedeling aan te spreek.<br />

Die atmosfeer van die aarde bestaan skynbaar uit konsentriese lae waarvan die eienskappe merkbaar<br />

verskil. Die atmosfeer is die kleinste van die reservoirs op aarde en is daarom mees vatbaar vir<br />

besoedeling. In hierdie afdeling word die struktuur en eienskappe van die atmosfeer beskryf. Die<br />

inligting behoort die nodige agtergrond te verskaf vir 'n latere bespreking van die chemiese prosesse<br />

in die atmosfeer.<br />

GEVALLESTUDIE: Aardverwarming<br />

Fokusvrae<br />

1. Wat is die kweekhuiseffek en wat is aardverwarming?<br />

2. Wat is die meganisme van aardverwarming?<br />

3. Watter getuienis is daar vir aardverwarming?<br />

4. Wat kan moontlike gevolge van aardverwarming wees?<br />

3.2. Die atmosfeer en lugbesoedeling<br />

3.2.1. Studiemateriaal<br />

(a) Air pollution, [Living in the environment], p. 467 – 490.<br />

(b) Verdere notas mag moontlik in die klas verskaf word.<br />

7


Chemie en die atmosfeer<br />

3.2.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

3.3. Klimaatontwrigting en osoonvernietiging<br />

3.3.1. Studiemateriaal<br />

(a) DVD: An inconvenient truth.<br />

(b) Climate disruption and ozone depletion, [Living in the environment], p. 491 – 527.<br />

3.3.2. Self-evaluering<br />

(a) Groepsbespreking oor die film "An inconvenient truth".<br />

(b) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

8


4. Chemie en die hidrosfeer<br />

TYD BENODIG Dit sal ongeveer 32 ure neem om hierdie leereenheid te voltooi.<br />

UITKOMSTE Beskryf die oorsake en gevolge van waterbesoedeling.<br />

Lys en beskryf kortliks die mees algemene siektes wat deur gekontamineerde<br />

water aan die mens oorgedra word.<br />

Bespreek die aard en oorsake van eutrofikasie.<br />

Bespreek hardnekkige organiese besoedelstowwe (persistent organic<br />

pollutants, POPs) in verhouding tot die hidrosfeer. (Hierdie verbindings is<br />

reeds bespreek in die toksikologie afdeling.)<br />

Bespreek die basiese metodes van watersuiwering en rioolverwerking.<br />

Bespreek die vorming en gevolge van suurmynwater.<br />

Stel wyses voor om waterbesoedeling aan te spreek.<br />

4.1. Inleiding<br />

Water is essensieel vir alle lewe op aarde. Toename in die bevolking van die aarde plaas groot druk<br />

op die beskikbare waterbronne en is besig om in 'n waterkrisis te ontaard. Daar is 'n groeiende<br />

konsensus dat die huidige watergebruikspatroon tot 'n waterskaarste asook 'n afname in waterkwaliteit<br />

sal lei. Hierdie faktore sal uiteindelik bepalend en beperkend wees vir toekomstige ekonomiese<br />

ontwikkeling, uitbreiding van kosproduksie en die verskaffing van basiese gesondheidsdienste en<br />

sanitêre dienste aan miljoene mense in ontwikkelende lande.<br />

4.2. Waterbesoedeling<br />

GEVALLESTUDIE: Die Carolina waterkrisis<br />

Fokusvrae<br />

1. Wat is die bron van die suur in Carolina se watertoevoer?<br />

2. Wat is die chemiese oorsprong van suurmynwater (AMD)?<br />

3. Wat is van die omgewingsprobleme wat deur suurmynwater veroorsaak word?<br />

4.2.1. Studiemateriaal<br />

(a) Water pollution, [Living in the environment], p. 528 – 556.<br />

4.2.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

9


Chemie en hidrosfeer<br />

4.3. Watersuiwering<br />

4.3.1. Studiemateriaal<br />

(a) VELDUITSTAPPIE: Daar word van studente vereis om deel te neem aan 'n velduitstappie na 'n<br />

watersuiweringsaanleg.<br />

(b) [Environmental chemistry], p. 601 – 660. Sien Bylaag 3.<br />

4.3.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

10


5. Chemie en die litosfeer<br />

TYD BENODIG Dit sal ongeveer 32 ure neem om hierdie leereenheid te voltooi.<br />

UITKOMSTE Verduidelik die terme "mineraal" en "gesteente" en verduidelik hoe hierdie<br />

entiteite deur middel van die rotssiklus verband hou.<br />

Onderskei tussen stollings-, sedimentêre- en metamorfiese gesteentes.<br />

Gee 'n oorsig van die Suid-Afrikaanse mynbedryf.<br />

Beskryf hoe ekonomies-lewensvatbare erts neerslae op aarde gevorm het.<br />

Illustreer en beskryf die algemene prosedure wat gebruik word om metale te<br />

ekstraheer.<br />

Beskryf die algemene chemiese beginsels wat gebruik word wanneer metale<br />

geëkstraheer word.<br />

Beskryf die basiese ekstraksieprosesse wat gebruik word tydens die<br />

ekstraksie van geselekteerde metale wat van belang is in die Suid-Afrikaanse<br />

konteks.<br />

Bespreek die voorkoms en gevolge van swaarmetaalbesoedeling.<br />

Stel wyses voor om besoedeling van die litosfeer aan te spreek.<br />

GEVALLESTUDIE: Die Sabierivier in 1922<br />

Fokusvrae<br />

1. Wat was die toestand van die Sabierivier in 1922?<br />

2. Hoe dra mynbou by tot omgewingsprobleme?<br />

3. Wat is die gevare van onwettige mynbou?<br />

5.1. Inleiding<br />

Hierdie afdeling bespreek aspekte van algemene geologie en mynbou met die klem op die Suid-<br />

Afrikaanse mynbedryf. Suid-Afrika is een van die belangrikste mynboulande in Afrika en in die wêreld<br />

gebaseer op die verskeidenheid en howeveelheid minerale wat gemyn word. Ons lang beskik oor die<br />

grootste reserwes van chroom, goud, vanadium, mangaan en PGMs in die wêreld. Die mynbou<br />

industrie kan verdeel word in vyf breë kategorieë – goud, PGMs, diamante, steenkool en vanadium. In<br />

totaal dra hierdie sektore ongeveer 6,5% bygedra tot die BBP in 2000. Die studiemateriaal behoort die<br />

student toe te rus met genoegsame agtergrondkennis om van die omgewingsuitdagings te begryp wat<br />

met mynbou geassosieer word. 'n Paar van hierdie probleme sal later in die studiemateriaal aandag<br />

kry.<br />

5.2. Geologie en mineraalbronne<br />

5.2.1. Studiemateriaal<br />

(a) Geology and non-renewable mineral resources, [Living in the environment], p. 346 – 369.<br />

11


Chemie en die litosfeer<br />

5.2.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

5.3. Basiese beginsels van mynbou en die ekstraksie van metale<br />

12


5.3.1. Studiemateriaal<br />

(a) Verslae van die Departement van Minerale en Energie (DME).<br />

(b) [Advanced Chemistry], p. 291 – 303, 518 – 524. Sien Bylaag 4.<br />

(c) [General chemistry], p. 996 – 1003. Sien Bylaag 5.<br />

5.3.2. Self-evaluering<br />

(a) Geselekteerde vrae gebaseer op die studiemateriaal sal in die klas verskaf word.<br />

Chemie en die litosfeer<br />

13


BYLAAG 1<br />

Manahan, SE, 2003, Toxicological chemistry and biochemistry, CRC Press, Boca Raton, p. 59 – 95.<br />

14


3.1 BIOCHEMISTRY<br />

CHAPTER 3<br />

Biochemistry<br />

Most people have had the experience of looking through a microscope at a single cell. It may<br />

have been an amoeba, alive and oozing about like a blob of jelly on the microscope slide, or a cell<br />

of bacteria, stained with a dye to make it show up more plainly. Or it may have been a beautiful<br />

cell of algae with its bright green chlorophyll. Even the simplest of these cells is capable of carrying<br />

out a thousand or more chemical reactions. These life processes fall under the heading of biochemistry,<br />

the branch of chemistry that deals with the chemical properties, composition, and<br />

biologically mediated processes of complex substances in living systems.<br />

Biochemical phenomena that occur in living organisms are extremely sophisticated. In the<br />

human body, complex metabolic processes break down a variety of food materials to simpler<br />

chemicals, yielding energy and the raw materials to build body constituents, such as muscle, blood,<br />

and brain tissue. Impressive as this may be, consider a humble microscopic cell of photosynthetic<br />

cyanobacteria only about a micrometer in size, which requires only a few simple inorganic chemicals<br />

and sunlight for its existence. This cell uses sunlight energy to convert carbon from CO2, hydrogen<br />

– 2– and oxygen from H2O, nitrogen from NO3, sulfur from SO4 , and phosphorus from inorganic<br />

phosphate into all the proteins, nucleic acids, carbohydrates, and other materials that it requires to<br />

exist and reproduce. Such a simple cell accomplishes what could not be done by human endeavors<br />

even in a vast chemical factory costing billions of dollars.<br />

Ultimately, most environmental pollutants and hazardous substances are of concern because of<br />

their effects on living organisms. The study of the adverse effects of substances on life processes<br />

requires some basic knowledge of biochemistry. Biochemistry is discussed in this chapter, with an<br />

emphasis on the aspects that are especially pertinent to environmentally hazardous and toxic<br />

substances, including cell membranes, deoxyribonucleic acid (DNA), and enzymes.<br />

Biochemical processes not only are profoundly influenced by chemical species in the environment,<br />

but they largely determine the nature of these species, their degradation, and even their<br />

syntheses, particularly in the aquatic and soil environments. The study of such phenomena forms<br />

the basis of environmental biochemistry. 1<br />

3.1.1 Biomolecules<br />

The biomolecules that constitute matter in living organisms are often polymers with molecular<br />

masses of the order of a million or even larger. As discussed later in this chapter, these biomolecules<br />

may be divided into the categories of carbohydrates, proteins, lipids, and nucleic acids. Proteins<br />

and nucleic acids consist of macromolecules, lipids are usually relatively small molecules, and<br />

carbohydrates range from relatively small sugar molecules to high-molecular-mass macromolecules,<br />

such as those in cellulose.<br />

15<br />

59


60 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Vacuole<br />

Ribosome<br />

Figure 3.1 Some major features of the eukaryotic cell in animals (left) and plants (right).<br />

The behavior of a substance in a biological system depends to a large extent upon whether the<br />

substance is hydrophilic (water-loving) or hydrophobic (water-hating). Some important toxic substances<br />

are hydrophobic, a characteristic that enables them to traverse cell membranes readily. Part<br />

of the detoxification process carried on by living organisms is to render such molecules hydrophilic,<br />

therefore water soluble and readily eliminated from the body.<br />

3.2 BIOCHEMISTRY AND THE CELL<br />

The focal point of biochemistry and biochemical aspects of toxicants is the cell, the basic<br />

building block of living systems where most life processes are carried. Bacteria, yeasts, and some<br />

algae consist of single cells. However, most living things are made up of many cells. In a more<br />

complicated organism the cells have different functions. Liver cells, muscle cells, brain cells, and<br />

skin cells in the human body are quite different from each other and do different things. Cells are<br />

divided into two major categories depending upon whether or not they have a nucleus: eukaryotic<br />

cells have a nucleus, and prokaryotic cells do not. Prokaryotic cells are found in single-celled<br />

bacteria. Eukaryotic cells compose organisms other than bacteria.<br />

3.2.1 Major Cell Features<br />

Golgi body<br />

Cell membrane<br />

Mitochondria<br />

Nucleus<br />

Lysosome<br />

Starch grain<br />

Mitochondria<br />

Chloroplast<br />

Cell wall<br />

Vacuole<br />

Figure 3.1 shows the major features of the eukaryotic cell, which is the basic structure in which<br />

biochemical processes occur in multicelled organisms. These features are as follows:<br />

• Cell membrane, which encloses the cell and regulates the passage of ions, nutrients, lipid-soluble<br />

(fat-soluble) substances, metabolic products, toxicants, and toxicant metabolites into and out of<br />

the cell interior because of its varying permeability for different substances. The cell membrane<br />

protects the contents of the cell from undesirable outside influences. Cell membranes are composed<br />

in part of phospholipids that are arranged with their hydrophilic (water-seeking) heads on the cell<br />

membrane surfaces and their hydrophobic (water-repelling) tails inside the membrane. Cell membranes<br />

contain bodies of proteins that are involved in the transport of some substances through<br />

the membrane. One reason the cell membrane is very important in toxicology and environmental<br />

biochemistry is because it regulates the passage of toxicants and their products into and out of the<br />

cell interior. Furthermore, when its membrane is damaged by toxic substances, a cell may not<br />

function properly and the organism may be harmed.<br />

• Cell nucleus, which acts as a sort of “control center” of the cell. It contains the genetic directions<br />

the cell needs to reproduce itself. The key substance in the nucleus is DNA. Chromosomes in the<br />

16


BIOCHEMISTRY 61<br />

cell nucleus are made up of combinations of DNA and proteins. Each chromosome stores a separate<br />

quantity of genetic information. Human cells contain 46 chromosomes. When DNA in the nucleus<br />

is damaged by foreign substances, various toxic effects, including mutations, cancer, birth defects,<br />

and defective immune system function may occur.<br />

• Cytoplasm, which fills the interior of the cell not occupied by the nucleus. Cytoplasm is further<br />

divided into a water-soluble proteinaceous filler called cytosol, in which are suspended bodies<br />

called cellular organelles, such as mitochondria or, in photosynthetic organisms, chloroplasts.<br />

• Mitochondria, “powerhouses” that mediate energy conversion and utilization in the cell. Mitochondria<br />

are sites in which food materials — carbohydrates, proteins, and fats — are broken down<br />

to yield carbon dioxide, water, and energy, which is then used by the cell for its energy needs.<br />

The best example of this is the oxidation of the sugar glucose, C 6H 12O 6:<br />

C 6H 12O 6 + 6O 2 → 6CO 2 + 6H 2O + energy<br />

This kind of process is called cellular respiration.<br />

• Ribosomes, which participate in protein synthesis.<br />

• Endoplasmic reticulum, which is involved in the metabolism of some toxicants by enzymatic<br />

processes.<br />

• Lysosome, a type of organelle that contains potent substances capable of digesting liquid food<br />

material. Such material enters the cell through a “dent” in the cell wall, which eventually becomes<br />

surrounded by cell material. This surrounded material is called a food vacuole. The vacuole merges<br />

with a lysosome, and the substances in the lysosome bring about digestion of the food material.<br />

The digestion process consists largely of hydrolysis reactions in which large, complicated food<br />

molecules are broken down into smaller units by the addition of water.<br />

• Golgi bodies, which occur in some types of cells. These are flattened bodies of material that serve<br />

to hold and release substances produced by the cells.<br />

• Cell walls of plant cells. These are strong structures that provide stiffness and strength. Cell walls<br />

are composed mostly of cellulose, which will be discussed later in this chapter.<br />

• Vacuoles inside plant cells that often contain materials dissolved in water.<br />

• Chloroplasts in plant cells that are involved in photosynthesis (the chemical process that uses energy<br />

from sunlight to convert carbon dioxide and water to organic matter). Photosynthesis occurs in these<br />

bodies. Food produced by photosynthesis is stored in the chloroplasts in the form of starch grains.<br />

3.3 PROTEINS<br />

Proteins are nitrogen-containing organic compounds that are the basic units of life systems.<br />

Cytoplasm, the jelly-like liquid filling the interior of cells, is made up largely of protein. Enzymes,<br />

which act as catalysts of life reactions, are made of proteins; they are discussed later in the chapter.<br />

Proteins are composed of amino acids (Figure 3.2) joined together in huge chains. Amino acids are<br />

organic compounds that contain the carboxylic acid group, – CO 2H, and the amino group, – NH 2. They<br />

are sort of a hybrid of carboxylic acids and amines (see Sections 1.8.1 and 1.8.2). Proteins are polymers,<br />

or macromolecules, of amino acids containing from approximately 40 to several thousand amino acid<br />

groups joined by peptide linkages. Smaller molecule amino acid polymers, containing only about 10<br />

to about 40 amino acids per molecule, are called polypeptides. A portion of the amino acid left after<br />

the elimination of H 2O during polymerization is called a residue. The amino acid sequence of these<br />

residues is designated by a series of three-letter abbreviations for the amino acid.<br />

Natural amino acids all have the following chemical group:<br />

H H<br />

N O<br />

R C<br />

H<br />

C<br />

17<br />

OH


62 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

H O<br />

H H O<br />

H H O<br />

H C C OH HO C C C OH H3C C C C OH<br />

NH2 H NH2 Glycine (gly) Serine (ser)<br />

OH NH2 H H O<br />

C C C OH<br />

H NH2 CH3 H H O<br />

H C C C C OH<br />

H H H O<br />

CH3 H<br />

H<br />

NH 3C C C C C OH<br />

2<br />

H CH3 NH2 Isoleucine (ile)*<br />

H H O<br />

O H H H O<br />

HS C C C OH<br />

H NH H3C 2<br />

S<br />

H<br />

C<br />

H<br />

H<br />

C<br />

H<br />

H O H2N C C C C C OH<br />

C C OH H H NH2 NH2 Methionine (met)*<br />

H<br />

H O<br />

3C H<br />

H H H H H O<br />

H<br />

C C C OH<br />

3N C C C C C C OH<br />

H3C H<br />

NH<br />

H H H NH2 2<br />

O H H O<br />

H2N C C C C OH<br />

H NH2 +<br />

H H O<br />

C C C OH<br />

H O<br />

H N NH<br />

NH2 H3C C C OH<br />

NH<br />

Phenylalanine (phe)*<br />

2<br />

Alanine (ala)<br />

H H<br />

H C C H O<br />

H C C C OH<br />

N<br />

H H<br />

H<br />

Cysteine (cys) Glutamine (gin)<br />

H H O<br />

HO C C C OH<br />

Valine (val)* H NH2 Lysine (lys)*<br />

H H O Tyrosine (tyr)<br />

C C C OH<br />

O H H O<br />

H NH<br />

HO C C C C OH<br />

N<br />

2<br />

H NH2 H<br />

Tryptophan (try)*<br />

+<br />

Threonine (thr)*<br />

H<br />

Histidine (his)<br />

Leucine (leu)*<br />

Proline (pro)<br />

Aspartic acid (asp)<br />

Asparagine (asn)<br />

O H H H O<br />

H H H H H O<br />

HO C C<br />

H<br />

C<br />

H<br />

C C OH<br />

NH2 H2N C N<br />

+<br />

NH2 C<br />

H<br />

C<br />

H<br />

C<br />

H<br />

C C OH<br />

NH2 Glutamic acid (glu)<br />

Arginine (arg)<br />

Figure 3.2 Amino acids that occur in proteins. Those marked with an asterisk cannot be synthesized by the<br />

human body and must come from dietary sources.<br />

In this structure the –NH 2 group is always bonded to the carbon next to the –CO 2H group. This is<br />

called the “alpha” location, so natural amino acids are alpha-amino acids. Other groups, designated as<br />

R, are attached to the basic alpha-amino acid structure. The R groups may be as simple as an atom of<br />

H found in glycine, or they may be as complicated as the structure of the R group in tryptophan:<br />

H H<br />

N O<br />

H C C<br />

H<br />

OH<br />

Zwitterion<br />

form<br />

Glycine<br />

18<br />

H<br />

H<br />

+<br />

N H O<br />

H C<br />

H<br />

C<br />

O –


BIOCHEMISTRY 63<br />

Alanine<br />

Figure 3.3 Condensation of alanine, leucine, and tyrosine to form a tripeptide consisting of three amino acids<br />

joined by peptide linkages (outlined by dashed lines).<br />

Table 3.1 Major Types of Proteins<br />

H<br />

H2N C C<br />

CH 3<br />

O<br />

OH<br />

H<br />

H2N C C<br />

CH 3<br />

+ H2N C C OH +<br />

H3C C CH3 Leucine H<br />

O<br />

H2N C C<br />

Tyrosine<br />

Type of Protein Example Function and Characteristics<br />

H<br />

Nutrient Casein (milk protein) Food source; people must have an adequate supply<br />

of nutrient protein with the right balance of amino<br />

acids for adequate nutrition<br />

Storage Ferritin Storage of iron in animal tissues<br />

Structural Collagen (tendons), keratin (hair) Structural and protective components in organisms<br />

Contractile Actin, myosin in muscle tissue Strong, fibrous proteins that can contract and cause<br />

movement to occur<br />

Transport Hemoglobin Transport inorganic and organic species across cell<br />

membranes, in blood, between organs<br />

Defense — Antibodies against foreign agents such as viruses<br />

produced by the immune system<br />

Regulatory Insulin, human growth hormone Regulate biochemical processes such as sugar<br />

metabolism or growth by binding to sites inside cells<br />

or on cell membranes<br />

Enzymes Acetylcholine esterase Catalysts of biochemical reactions (see Section 3.6)<br />

As shown in Figure 3.2, there are 20 common amino acids in proteins. These are shown with<br />

uncharged – NH 2 and – CO 2H groups. Actually, these functional groups exist in the charged zwitterion<br />

form, as shown for glycine above.<br />

Amino acids in proteins are joined together in a specific way. These bonds constitute the peptide<br />

linkage. The formation of peptide linkages is a condensation process involving the loss of water.<br />

For example, consider the condensation of alanine, leucine, and tyrosine shown in Figure 3.3. When<br />

these three amino acids join together, two water molecules are eliminated. The product is a tripeptide<br />

since there are three amino acids involved. The amino acids in proteins are linked as shown for<br />

this tripeptide, except that many more monomeric amino acid groups are involved.<br />

Proteins may be divided into several major types that have widely varying functions. These are<br />

listed in Table 3.1.<br />

O<br />

H C H<br />

H<br />

N C<br />

H C<br />

H 3C<br />

H O<br />

C<br />

H<br />

19<br />

C<br />

H<br />

CH 3<br />

H<br />

N<br />

H<br />

O<br />

C C<br />

H C H<br />

OH<br />

H<br />

O<br />

H C H<br />

OH<br />

OH<br />

OH


64 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

3.3.1 Protein Structure<br />

The order of amino acids in protein molecules, and the resulting three-dimensional structures<br />

that form, provide an enormous variety of possibilities for protein structure. This is what makes<br />

life so diverse. Proteins have primary, secondary, tertiary, and quaternary structures. The structures<br />

of protein molecules determine the behavior of proteins in crucial areas such as the processes by<br />

which the body’s immune system recognizes substances that are foreign to the body. Proteinaceous<br />

enzymes depend on their structures for the very specific functions of the enzymes.<br />

The order of amino acids in the protein molecule determines its primary structure. Secondary<br />

protein structures result from the folding of polypeptide protein chains to produce a maximum<br />

number of hydrogen bonds between peptide linkages:<br />

Hydrogen bonds<br />

C O H N Illustration of hydrogen bonds between<br />

N and O atoms in peptide linkages, which<br />

N H O C<br />

constitutes protein secondary structures<br />

Hydrogen bonds<br />

Further folding of the protein molecules held in place by attractive forces between amino acid side<br />

chains gives proteins a secondary structure, which is determined by the nature of the amino acid<br />

R groups. Small R groups enable protein molecules to be hydrogen-bonded together in a parallel<br />

arrangement, whereas large R groups produce a spiral form known as an alpha-helix.<br />

Tertiary structures are formed by the twisting of alpha-helices into specific shapes. They are<br />

produced and held in place by the interactions of amino side chains on the amino acid residues<br />

constituting the protein macromolecules. Tertiary protein structure is very important in the processes<br />

by which enzymes identify specific proteins and other molecules upon which they act. It is also<br />

involved with the action of antibodies in blood, which recognize foreign proteins by their shape<br />

and react to them. This is what happens in the phenomenon of disease immunity, where antibodies<br />

in blood recognize specific proteins from viruses or bacteria and reject them.<br />

Two or more protein molecules consisting of separate polypeptide chains may be further<br />

attracted to each other to produce a quaternary structure.<br />

Some proteins are fibrous proteins, which occur in skin, hair, wool, feathers, silk, and tendons.<br />

The molecules in these proteins are long and threadlike and are laid out parallel in bundles. Fibrous<br />

proteins are quite tough and do not dissolve in water.<br />

An interesting fibrous protein is keratin, which is found in hair. The cross-linking bonds between<br />

protein molecules in keratin are –S–S– bonds formed from two HS– groups in two molecules of<br />

the amino acid cysteine. These bonds largely hold hair in place, thus keeping it curly or straight.<br />

A “permanent” consists of breaking the bonds chemically, setting the hair as desired, and then<br />

reforming the cross-links to hold the desired shape.<br />

Aside from fibrous protein, the other major type of protein form is the globular protein. These<br />

proteins are in the shape of balls and oblongs. Globular proteins are relatively soluble in water. A<br />

typical globular protein is hemoglobin, the oxygen-carrying protein in red blood cells. Enzymes<br />

are generally globular proteins.<br />

20


BIOCHEMISTRY 65<br />

3.3.2 Denaturation of Proteins<br />

Secondary, tertiary, and quaternary protein structures are easily changed by a process called<br />

denaturation. These changes can be quite damaging. Heating, exposure to acids or bases, and even<br />

violent physical action can cause denaturation to occur. The albumin protein in egg white is<br />

denatured by heating so that it forms a semisolid mass. Almost the same thing is accomplished by<br />

the violent physical action of an egg beater in the preparation of meringue. Heavy metal poisons<br />

such as lead and cadmium change the structures of proteins by binding to functional groups on the<br />

protein surface.<br />

3.4 CARBOHYDRATES<br />

Carbohydrates have the approximate simple formula CH 2O and include a diverse range of<br />

substances composed of simple sugars such as glucose:<br />

CH2OH H<br />

C<br />

HO<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

H<br />

C<br />

OH<br />

H OH<br />

Glucose molecule<br />

High-molecular-mass polysaccharides, such as starch and glycogen (animal starch), are biopolymers<br />

of simple sugars.<br />

Photosynthesis in a plant cell converts the energy from sunlight to chemical energy in a<br />

carbohydrate, C 6H 12O 6. This carbohydrate may be transferred to some other part of the plant for<br />

use as an energy source. It may be converted to a water-insoluble carbohydrate for storage until it<br />

is needed for energy. Or it may be transformed to cell wall material and become part of the structure<br />

of the plant. If the plant is eaten by an animal, the carbohydrate is used for energy by the animal.<br />

The simplest carbohydrates are the monosaccharides. These are also called simple sugars.<br />

Because they have six carbon atoms, simple sugars are sometimes called hexoses. Glucose (formula<br />

shown above) is the most common simple sugar involved in cell processes. Other simple sugars<br />

with the same formula but somewhat different structures are fructose, mannose, and galactose.<br />

These must be changed to glucose before they can be used in a cell. Because of its use for energy<br />

in body processes, glucose is found in the blood. Normal levels are from 65 to 110 mg of glucose<br />

per 100 ml of blood. Higher levels may indicate diabetes.<br />

Units of two monosaccharides make up several very important sugars known as disaccharides.<br />

When two molecules of monosaccharides join together to form a disaccharide,<br />

C 6H 12O 6 + C 6H 12O 6 → C 12H 22O 11 + H 2O (3.4.1)<br />

a molecule of water is lost. Recall that proteins are also formed from smaller amino acid molecules<br />

by condensation reactions involving the loss of water molecules. Disaccharides include sucrose (cane<br />

sugar used as a sweetener), lactose (milk sugar), and maltose (a product of the breakdown of starch).<br />

Polysaccharides consist of many simple sugar units hooked together. One of the most important<br />

polysaccharides is starch, which is produced by plants for food storage. Animals produce a related<br />

material called glycogen. The chemical formula of starch is (C 6H 10O 5) n, where n may represent a<br />

number as high as several hundred. What this means is that the very large starch molecule consists<br />

21


66 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

CH2OH CH2OH CH2OH H<br />

C<br />

O<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

H H<br />

C C<br />

O<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

H H<br />

C C<br />

O<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

H<br />

C<br />

O<br />

H OH H OH H OH<br />

Figure 3.4 Part of a starch molecule showing units of C 6H 10O 5 condensed together.<br />

CH2OH H OH<br />

H<br />

C<br />

O<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

O<br />

C C<br />

H H<br />

C<br />

OH<br />

H<br />

C<br />

C<br />

H<br />

O<br />

H<br />

C<br />

O<br />

H OH CH2OH Figure 3.5 Part of the structure of cellulose.<br />

of many units of C 6H 10O 5 joined together. For example, if n is 100, there are 6 times 100 carbon<br />

atoms, 10 times 100 hydrogen atoms, and 5 times 100 oxygen atoms in the molecule. Its chemical<br />

formula is C 600H 1000O 500. The atoms in a starch molecule are actually present as linked rings,<br />

represented by the structure shown in Figure 3.4. Starch occurs in many foods, such as bread and<br />

cereals. It is readily digested by animals, including humans.<br />

Cellulose is a polysaccharide that is also made up of C 6H 10O 5 units. Molecules of cellulose are<br />

huge, with molecular weights of around 400,000. The cellulose structure (Figure 3.5) is similar to<br />

that of starch. Cellulose is produced by plants and forms the structural material of plant cell walls.<br />

Wood is about 60% cellulose, and cotton contains over 90% of this material. Fibers of cellulose<br />

are extracted from wood and pressed together to make paper.<br />

Humans and most other animals cannot digest cellulose. Ruminant animals (cattle, sheep, goats,<br />

moose) have bacteria in their stomachs that break down cellulose into products that can be used<br />

by the animal. Chemical processes are available to convert cellulose to simple sugars by the reaction<br />

(C 6H 10O 5) n + nH 2O → nC 6H 12O 6<br />

cellulose glucose<br />

(3.4.2)<br />

where n may be 2000 to 3000. This involves breaking the linkages between units of C 6H 10O 5 by<br />

adding a molecule of H 2O at each linkage, a hydrolysis reaction. Large amounts of cellulose from<br />

wood, sugar cane, and agricultural products go to waste each year. The hydrolysis of cellulose<br />

enables these products to be converted to sugars, which can be fed to animals.<br />

Carbohydrate groups are attached to protein molecules in a special class of materials called<br />

glycoproteins. Collagen is a crucial glycoprotein that provides structural integrity to body parts.<br />

It is a major constituent of skin, bones, tendons, and cartilage.<br />

3.5 LIPIDS<br />

Lipids are substances that can be extracted from plant or animal matter by organic solvents,<br />

such as chloroform, diethyl ether, or toluene (Figure 3.6). Whereas carbohydrates and proteins are<br />

22<br />

H<br />

C<br />

CH2OH C O<br />

H<br />

OH<br />

C<br />

H<br />

C<br />

O<br />

C<br />

H<br />

H OH


BIOCHEMISTRY 67<br />

Cooling<br />

water in<br />

Rising solvent vapor<br />

Porous thimble<br />

containing sample<br />

Condenser<br />

Boiling solvent<br />

Cooling<br />

water out<br />

Condensed solvent<br />

Siphon back to<br />

solvent reservoir<br />

Heating mantle<br />

Figure 3.6 Lipids are extracted from some biological materials with a soxhelet extractor (above). The solvent<br />

is vaporized in the distillation flask by the heating mantle, rises through one of the exterior tubes<br />

to the condenser, and is cooled to form a liquid. The liquid drops onto the porous thimble containing<br />

the sample. Siphon action periodically drains the solvent back into the distillation flask. The extracted<br />

lipid collects as a solution in the solvent in the flask.<br />

characterized predominately by the monomers (monosaccharides and amino acids) from which<br />

they are composed, lipids are defined essentially by their physical characteristic of organophilicity.<br />

The most common lipids are fats and oils composed of triglycerides formed from alcohol glycerol,<br />

CH 2(OH)CH(OH)CH 2(OH), and a long-chain fatty acid such as stearic acid, CH 3(CH 2) 16C(O)OH<br />

(Figure 3.7). Numerous other biological materials, including waxes, cholesterol, and some vitamins<br />

and hormones, are classified as lipids. Common foods, such as butter and salad oils, are lipids.<br />

Long-chain fatty acids, such as stearic acid, are also organic soluble and are classified as lipids.<br />

Lipids are toxicologically important for several reasons. Some toxic substances interfere with<br />

lipid metabolism, leading to detrimental accumulation of lipids. Many toxic organic compounds<br />

are poorly soluble in water, but are lipid soluble, so that bodies of lipids in organisms serve to<br />

dissolve and store toxicants.<br />

An important class of lipids consists of phosphoglycerides (glycerophosphatides). These compounds<br />

may be regarded as triglycerides in which one of the acids bonded to glycerol is ortho-<br />

23


68 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Figure 3.7 General formula of triglycerides, which make up fats and oils. The R group is from a fatty acid and<br />

is a hydrocarbon chain, such as –(CH 2) 16CH 3.<br />

phosphoric acid. These lipids are especially important because they are essential constituents of<br />

cell membranes. These membranes consist of bilayers in which the hydrophilic phosphate ends of<br />

the molecules are on the outside of the membrane and the hydrophobic “tails” of the molecules<br />

are on the inside.<br />

Waxes are also esters of fatty acids. However, the alcohol in a wax is not glycerol; it is often<br />

a very long chain alcohol. For example, one of the main compounds in beeswax is myricyl palmitate,<br />

H O<br />

(C30H61 ) C O<br />

H<br />

C<br />

(C 15 H 31 )<br />

Alcohol portion Fatty acid portion<br />

of ester of ester<br />

in which the alcohol portion of the ester has a very large hydrocarbon chain. Waxes are produced<br />

by both plants and animals, largely as protective coatings. Waxes are found in a number of common<br />

products. Lanolin is one of these. It is the “grease” in sheep’s wool. When mixed with oils and<br />

water, it forms stable colloidal emulsions consisting of extremely small oil droplets suspended in<br />

water. This makes lanolin useful for skin creams and pharmaceutical ointments. Carnauba wax<br />

occurs as a coating on the leaves of some Brazilian palm trees. Spermaceti wax is composed largely<br />

of cetyl palmitate,<br />

(C 15 H 31 )<br />

H O<br />

C O C<br />

H<br />

(C 15 H 31 )<br />

Cetyl palmitate<br />

which is extracted from the blubber of the sperm whale. It is very useful in some cosmetics and<br />

pharmaceutical preparations.<br />

Steroids are lipids found in living systems that all have the ring system shown in Figure 3.8<br />

for cholesterol. Steroids occur in bile salts, which are produced by the liver and then secreted into<br />

the intestines. Their breakdown products give feces its characteristic color. Bile salts act on fats in<br />

the intestine. They suspend very tiny fat droplets in the form of colloidal emulsions. This enables<br />

the fats to be broken down chemically and digested.<br />

Some steroids are hormones. Hormones act as “messengers” from one part of the body to<br />

another. As such, they start and stop a number of body functions. Male and female sex hormones<br />

are examples of steroid hormones. Hormones are given off by glands in the body called endocrine<br />

glands. The locations of the important endocrine glands are shown in Figure 3.9.<br />

24


BIOCHEMISTRY 69<br />

HO<br />

H 3 C<br />

H 3 C<br />

H 3 C<br />

Figure 3.8 Steroids are characterized by the ring structure shown above for cholesterol.<br />

H<br />

C<br />

Pituitary<br />

Parathyroid<br />

Ovaries<br />

(female)<br />

Figure 3.9 Locations of important endocrine glands.<br />

CH 2 CH 2 CH 2 C H<br />

3.6 ENZYMES<br />

Catalysts are substances that speed up a chemical reaction without themselves being consumed<br />

in the reaction. The most sophisticated catalysts of all are those found in living systems. They bring<br />

about reactions that could not be performed at all, or only with great difficulty, outside a living<br />

organism. These catalysts are called enzymes. In addition to speeding up reactions by as much as<br />

10- to a 100 million-fold, enzymes are extremely selective in the reactions they promote.<br />

Enzymes are proteinaceous substances with highly specific structures that interact with particular<br />

substances or classes of substances called substrates. Enzymes act as catalysts to enable<br />

biochemical reactions to occur, after which they are regenerated intact to take part in additional<br />

reactions. The extremely high specificity with which enzymes interact with substrates results from<br />

their “lock and key” action, based on the unique shapes of enzymes, as illustrated in Figure 3.10.<br />

This illustration shows that an enzyme “recognizes” a particular substrate by its molecular<br />

structure and binds to it to produce an enzyme–substrate complex. This complex then breaks apart<br />

to form one or more products different from the original enzyme, regenerating the unchanged<br />

enzyme, which is then available to catalyze additional reactions. The basic process for an enzyme<br />

reaction is, therefore,<br />

CH 3<br />

CH 3<br />

Cholesterol, a typical steroid<br />

Thymus<br />

Adrenal<br />

25<br />

Thyroid<br />

Testes<br />

(male)


70 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Enzyme<br />

+<br />

Substrate<br />

+<br />

Enzyme-substrate complex<br />

Products Regenerated enzyme<br />

Figure 3.10 Representation of the “lock and key” mode of enzyme action, which enables the very high specificity<br />

of enzyme-catalyzed reactions.<br />

enzyme + substrate ⎯→ ⎯⎯←⎯<br />

enzyme–substrate complex ⎯→ ⎯⎯←⎯<br />

enzyme + product (3.6.1)<br />

Several important things should be noted about this reaction. As shown in Figure 3.10, an enzyme<br />

acts on a specific substrate to form an enzyme–substrate complex because of the fit between their<br />

structures. As a result, something happens to the substrate molecule. For example, it might be split<br />

in two at a particular location. Then the enzyme–substrate complex comes apart, yielding the<br />

enzyme and products. The enzyme is not changed in the reaction and is now free to react again.<br />

Note that the arrows in the formula for enzyme reaction point both ways. This means that the<br />

reaction is reversible. An enzyme–substrate complex can simply go back to the enzyme and the<br />

substrate. The products of an enzymatic reaction can react with the enzyme to form the enzyme–substrate<br />

complex again. It, in turn, may again form the enzyme and the substrate. Therefore, the same<br />

enzyme may act to cause a reaction to go either way.<br />

Some enzymes cannot function by themselves. In order to work, they must first be attached to<br />

coenzymes. Coenzymes normally are not protein materials. Some of the vitamins are important<br />

coenzymes.<br />

Enzymes are named for what they do. For example, the enzyme given off by the stomach,<br />

which splits proteins as part of the digestion process, is called gastric proteinase. The “gastric”<br />

part of the name refers to the enzyme’s origin in the stomach. “Proteinase” denotes that it splits<br />

up protein molecules. The common name for this enzyme is pepsin. Similarly, the enzyme produced<br />

by the pancreas that breaks down fats (lipids) is called pancreatic lipase. Its common name is<br />

26


BIOCHEMISTRY 71<br />

steapsin. In general, lipase enzymes cause lipid triglycerides to dissociate and form glycerol and<br />

fatty acids.<br />

The enzymes mentioned above are hydrolyzing enzymes, which bring about the breakdown<br />

of high-molecular-weight biological compounds by the addition of water. This is one of the most<br />

important reactions involved in digestion. The three main classes of energy-yielding foods that<br />

animals eat are carbohydrates, proteins, and fats. Recall that the higher carbohydrates that humans<br />

eat are largely disaccharides (sucrose, or table sugar) and polysaccharides (starch). These are formed<br />

by the joining together of units of simple sugars, C 6H 12O 6, with the elimination of an H 2O molecule<br />

at the linkage where they join. Proteins are formed by the condensation of amino acids, again with<br />

the elimination of a water molecule at each linkage. Fats are esters that are produced when glycerol<br />

and fatty acids link together. A water molecule is lost for each of these linkages when a protein,<br />

fat, or carbohydrate is synthesized. In order for these substances to be used as a food source, the<br />

reverse process must occur to break down large, complicated molecules of protein, fat, or carbohydrate<br />

to simple, soluble substances that can penetrate a cell membrane and take part in chemical<br />

processes in the cell. This reverse process is accomplished by hydrolyzing enzymes.<br />

Biological compounds with long chains of carbon atoms are broken down into molecules with<br />

shorter chains by the breaking of carbon–carbon bonds. This commonly occurs by the elimination<br />

of –CO 2H groups from carboxylic acids. For example, pyruvic decarboxylase enzyme acts upon<br />

pyruvic acid,<br />

H<br />

H O O<br />

H O<br />

C C C OH<br />

Pyruvate<br />

decarboxylase H C C H + CO2<br />

H<br />

H<br />

Pyruvic acid Acetaldehyde<br />

(3.6.2)<br />

to split off CO 2 and produce a compound with one less carbon. It is by such carbon-by-carbon<br />

breakdown reactions that long-chain compounds are eventually degraded to CO 2 in the body, or<br />

that long-chain hydrocarbons undergo biodegradation by the action of spill bacteria on spilled<br />

petroleum. Oxidation and reduction are the major reactions for the exchange of energy in living<br />

systems. Cellular respiration, discussed in Section 3.2, is an oxidation reaction in which a carbohydrate,<br />

C 6H 12O 6, is broken down to carbon dioxide and water with the release of energy:<br />

C 6H 12O 6 + 6O 2 → 6CO 2 + 6H 2O + energy (3.6.3)<br />

Actually, such an overall reaction occurs in living systems by a complicated series of individual<br />

steps. Some of these steps involve oxidation. The enzymes that bring about oxidation in the presence<br />

of free O 2 are called oxidases. In general, biological oxidation–reduction reactions are catalyzed<br />

by oxidoreductase enzymes.<br />

In addition to the types of enzymes discussed above, there are many enzymes that perform<br />

miscellaneous duties in living systems. Typical of these are isomerases, which form isomers of<br />

particular compounds. For example, there are several simple sugars with the formula C 6H 12O 6.<br />

However, only glucose can be used directly for cell processes. The other isomers are converted to<br />

glucose by the action of isomerases. Transferase enzymes move chemical groups from one<br />

molecule to another, lyase enzymes remove chemical groups without hydrolysis and participate in<br />

the formation of C=C bonds or addition of species to such bonds, and ligase enzymes work in<br />

conjunction with adenosine triphosphate (ATP), a high-energy molecule that plays a crucial role<br />

in energy-yielding, glucose-oxidizing metabolic processes, to link molecules together with the<br />

formation of bonds such as carbon–carbon or carbon–sulfur bonds.<br />

27


72 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Enzyme action may be affected by many different things. Enzymes require a certain hydrogen<br />

ion concentration to function best. For example, gastric proteinase requires the acid environment<br />

of the stomach to work well. When it passes into the much less acidic intestines, it stops working.<br />

This prevents damage to the intestine walls, which would occur if the enzyme tried to digest them.<br />

Temperature is critical. Not surprisingly, the enzymes in the human body work best at around<br />

98.6°F (37°C), which is the normal body temperature. Heating these enzymes to around 140°F<br />

permanently destroys them. Some bacteria that thrive in hot springs have enzymes that work best<br />

at relatively high temperatures. Other “cold-seeking” bacteria have enzymes adapted to near the<br />

freezing point of water.<br />

One of the greatest concerns regarding the effects of surroundings on enzymes is the influence<br />

of toxic substances. A major mechanism of toxicity is the alteration or destruction of enzymes by<br />

agents such as cyanide, heavy metals, or organic compounds, such as insecticidal parathion. An<br />

enzyme that has been destroyed obviously cannot perform its designated function, whereas one<br />

that has been altered either may not function at all or may act improperly. Toxicants can affect<br />

enzymes in several ways. Parathion, for example, bonds covalently to the nerve enzyme acetylcholinesterase,<br />

which can then no longer serve to stop nerve impulses. Heavy metals tend to bind to<br />

sulfur atoms in enzymes (such as sulfur from the amino acid cysteine, shown in Figure 3.2), thereby<br />

altering the shape and function of the enzyme. Enzymes are denatured by some poisons, causing<br />

them to “unravel” so that the enzyme no longer has its crucial specific shape.<br />

3.7 NUCLEIC ACIDS<br />

The essence of life is contained in deoxyribonucleic acid (DNA), which stays in the cell<br />

nucleus, and ribonucleic acid (RNA), which functions in the cell cytoplasm. These substances,<br />

which are known collectively as nucleic acids, store and pass on essential genetic information that<br />

controls reproduction and protein synthesis.<br />

The structural formulas of the monomeric constituents of nucleic acids are given in Figure 3.11.<br />

These are pyrimidine or purine nitrogen-containing bases, two sugars, and phosphate. DNA molecules<br />

are made up of the nitrogen-containing bases adenine, guanine, cytosine, and thymine;<br />

phosphoric acid (H 3PO 4); and the simple sugar 2-deoxy-β-D-ribofuranose (commonly called deoxyribose).<br />

RNA molecules are composed of the nitrogen-containing bases adenine, guanine, cytosine,<br />

and uracil; phosphoric acid (H 3PO 4); and the simple sugar β-D-ribofuranose (ribose).<br />

The formation of nucleic acid polymers from their monomeric constituents may be viewed as<br />

the following steps.<br />

• Monosaccharide (simple sugar) + cyclic nitrogenous base yields nucleoside:<br />

HO<br />

NH2 H<br />

N<br />

C<br />

C<br />

H<br />

O<br />

C C<br />

N H<br />

CH2 C<br />

H<br />

C C<br />

C<br />

H<br />

O<br />

H<br />

H<br />

Deoxyctidine formed by the<br />

dimerization of cytosine and<br />

deoxyribose with the elimination<br />

of a molecule of H2 O.<br />

HO H<br />

28


BIOCHEMISTRY 73<br />

HO<br />

H<br />

O<br />

N<br />

O<br />

C<br />

C<br />

C N C<br />

H<br />

Thymine (T)<br />

Occur only in DNA<br />

CH 3<br />

H<br />

CH2 C<br />

H<br />

C C<br />

C<br />

H<br />

O OH<br />

H<br />

H<br />

HO H<br />

2-Deoxy-β-Dribofuranose<br />

NH2 H<br />

N<br />

C<br />

C<br />

H<br />

O<br />

C C<br />

N H<br />

N C<br />

NH2 H<br />

N<br />

C<br />

C<br />

C<br />

N<br />

N<br />

C H<br />

H<br />

N C<br />

O<br />

H<br />

H2N N<br />

C<br />

C<br />

C<br />

N<br />

N<br />

C H<br />

H<br />

P<br />

O- O- - H<br />

Cytosine (C)<br />

Adenine (A)<br />

Guanine(G)<br />

O<br />

O Phosphate<br />

Occur in both DNA<br />

and RNA<br />

Figure 3.11 Constituents of DNA (enclosed by ----) and of RNA (enclosed by |||||).<br />

• Nucleoside + phosphate yields phosphate ester nucleotide.<br />

- O<br />

P<br />

O -<br />

O<br />

O<br />

NH 2<br />

HO<br />

H<br />

O<br />

N<br />

O<br />

C<br />

C<br />

C N C<br />

H<br />

Uracil (U)<br />

Occur only in RNA<br />

H<br />

H<br />

CH2 C<br />

H<br />

C C<br />

C<br />

H<br />

O OH<br />

H<br />

H<br />

HO OH<br />

β-D-Ribofuranose<br />

H<br />

N<br />

C<br />

C<br />

H<br />

O<br />

C C<br />

N H<br />

CH2 C<br />

H<br />

C C<br />

C Nucleotide formed by the<br />

bonding of a phosphate<br />

group to deoxyctidine<br />

H<br />

O O<br />

H<br />

H<br />

HO H<br />

29


74 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

• Polymerized nucleotide yields nucleic acid, as shown by the structure below. In the nucleic acid<br />

the phosphate negative charges are neutralized by metal cations (such as Mg 2+ ) or positively charged<br />

proteins (histones).<br />

O<br />

P<br />

O -<br />

O<br />

O<br />

CH 2<br />

NH 2<br />

H<br />

N<br />

C<br />

C<br />

H<br />

O<br />

C C<br />

N H<br />

O<br />

C<br />

H<br />

C C<br />

C<br />

H H<br />

H<br />

O H<br />

P<br />

O -<br />

N<br />

H C<br />

O O CH2 C<br />

H<br />

H<br />

C<br />

O<br />

H<br />

C<br />

N<br />

C<br />

H<br />

O H<br />

Segment of the DNA polymer<br />

showing linkage of two nucleotides<br />

Molecules of DNA are huge, with molecular weights of greater than 1 billion. Molecules of<br />

RNA are also quite large. The structure of DNA is that of the famed double helix. It was figured<br />

out in 1953 by James D. Watson, an American scientist, and Francis Crick, a British scientist. They<br />

received the Nobel Prize for this scientific milestone in 1962. This model visualizes DNA as a socalled<br />

double α-helix structure of oppositely wound polymeric strands held together by hydrogen<br />

bonds between opposing pyrimidine and purine groups. As a result, DNA has both a primary and<br />

a secondary structure; the former is due to the sequence of nucleotides in the individual strands of<br />

DNA, and the latter results from the α-helix interaction of the two strands. In the secondary structure<br />

of DNA, only cytosine can be opposite guanine and only thymine can be opposite adenine and<br />

vice versa. Basically, the structure of DNA is that of two spiral ribbons “counterwound” around<br />

each other, as illustrated in Figure 3.12. The two strands of DNA are complementary. This means<br />

that a particular portion of one strand fits like a key in a lock with the corresponding portion of<br />

another strand. If the two strands are pulled apart, each manufactures a new complementary strand,<br />

so that two copies of the original double helix result. This occurs during cell reproduction.<br />

The molecule of DNA is like a coded message. This “message,” the genetic information<br />

contained in and transmitted by nucleic acids, depends on the sequence of bases from which they<br />

are composed. It is somewhat like the message sent by telegraph, which consists only of dots,<br />

dashes, and spaces in between. The key aspect of DNA structure that enables storage and replication<br />

of this information is the famed double helix structure of DNA mentioned above.<br />

Portions of the DNA double helix may unravel, and one of the strands of DNA may produce<br />

a strand of RNA. This substance then goes from the cell nucleus out into the cell and regulates the<br />

synthesis of new protein. In this way, DNA regulates the function of the cell and acts to control<br />

life processes.<br />

30<br />

C<br />

C<br />

NH 2<br />

N C<br />

C<br />

N<br />

H


BIOCHEMISTRY 75<br />

3.7.1 Nucleic Acids in Protein Synthesis<br />

When a new cell is formed, the DNA in its nucleus must be accurately reproduced from the<br />

parent cell. Life processes are absolutely dependent upon accurate protein synthesis, as regulated<br />

by cell DNA. The DNA in a single cell must be capable of directing the synthesis of up to 3000<br />

or even more different proteins. The directions for the synthesis of a single protein are contained<br />

in a segment of DNA called a gene. The process of transmitting information from DNA to a newly<br />

synthesized protein involves the following steps:<br />

• The DNA undergoes replication. This process involves separation of a segment of the double<br />

helix into separate single strands, which then replicate such that guanine is opposite cytosine (and<br />

vice versa) and adenine is opposite thymine (and vice versa). This process continues until a<br />

complete copy of the DNA molecule has been produced.<br />

• The newly replicated DNA produces messenger RNA (mRNA), a complement of the single strand<br />

of DNA, by a process called transcription.<br />

• A new protein is synthesized using mRNA as a template to determine the order of amino acids in<br />

a process called translation.<br />

3.7.2 Modified DNA<br />

A T<br />

C G<br />

T A<br />

C G<br />

G<br />

A<br />

G<br />

Figure 3.12 Representation of the double helix structure of DNA showing the allowed base pairs held together<br />

by hydrogen bonding between the phosphate–sugar polymer “backbones” of the two strands of<br />

DNA. The letters stand for adenine (A), cytosine (C), guanine (G), and thymine (T). The dashed<br />

lines represent hydrogen bonds.<br />

DNA molecules may be modified by the unintentional addition or deletion of nucleotides or<br />

by substituting one nucleotide for another. The result is a mutation that is transmittable to offspring.<br />

Mutations can be induced by chemical substances. This is a major concern from a toxicological<br />

viewpoint because of the detrimental effects of many mutations and because substances that cause<br />

mutations often cause cancer as well. DNA malfunction may result in birth defects, and the failure<br />

to control cell reproduction results in cancer. Radiation from x-rays and radioactivity also disrupts<br />

DNA and may cause mutation.<br />

C<br />

T<br />

C<br />

T A<br />

C G<br />

31


76 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

3.8 RECOMBINANT DNA AND GENETIC ENGINEERING<br />

As noted above, segments of DNA contain information for the specific syntheses of particular<br />

proteins. Within the last two decades it has become possible to transfer this information between<br />

organisms by means of recombinant DNA technology, which has resulted in a new industry based<br />

on genetic engineering. Most often the recipient organisms are bacteria, which can be reproduced<br />

(cloned) over many orders of magnitude from a cell that has acquired the desired qualities.<br />

Therefore, to synthesize a particular substance, such as human insulin or growth hormone, the<br />

required genetic information can be transferred from a human source to bacterial cells, which then<br />

produce the substance as part of their metabolic processes.<br />

The first step in recombinant DNA gene manipulation is to lyze (open up) a donor cell to<br />

remove needed DNA material by using enzyme action to cut the sought-after genes from the donor<br />

DNA chain. These are next spliced into small DNA molecules. These molecules, called cloning<br />

vehicles, are capable of penetrating the host cell and becoming incorporated into its genetic material.<br />

The modified host cell is then reproduced many times and carries out the desired biosynthesis.<br />

Early concerns about the potential of genetic engineering to produce “monster organisms” or<br />

new and horrible diseases have been largely allayed, although caution is still required with this<br />

technology. In the environmental area, genetic engineering offers some hope for the production of<br />

bacteria engineered to safely destroy troublesome wastes and to produce biological substitutes for<br />

environmentally damaging synthetic pesticides.<br />

3.9 METABOLIC PROCESSES<br />

Biochemical processes that involve the alteration of biomolecules fall under the category of<br />

metabolism. Metabolic processes may be divided into the two major categories of anabolism<br />

(synthesis) and catabolism (degradation of substances). An organism may use metabolic processes<br />

to yield energy or to modify the constituents of biomolecules.<br />

3.9.1 Energy-Yielding Processes<br />

Organisms can gain energy by the following three processes:<br />

• Respiration, in which organic compounds undergo catabolism that requires molecular oxygen<br />

(aerobic respiration) or that occurs in the absence of molecular oxygen (anaerobic respiration).<br />

Aerobic respiration uses the Krebs cycle to obtain energy from the following reaction:<br />

C 6H 12O 6 + 6O 2 → 6CO 2 + 6H 2O + energy<br />

• About half of the energy released is converted to short-term stored chemical energy, particularly<br />

through the synthesis of ATP nucleoside. For longer-term energy storage, glycogen or starch<br />

polysaccharides are synthesized, and for still longer term energy storage, lipids (fats) are generated<br />

and retained by the organism.<br />

• Fermentation, which differs from respiration in not having an electron transport chain. Yeasts<br />

produce ethanol from sugars by fermentation:<br />

C 6H 12O 6 → 2CO 2 + 2C 2H 5OH<br />

• Photosynthesis, in which light energy captured by plant and algal chloroplasts are used to synthesize<br />

sugars from carbon dioxide and water:<br />

32


BIOCHEMISTRY 77<br />

6CO 2 + 6H 2O + hν → C 6H 12O 6 + 6O 2<br />

Plants cannot always get the energy that they need from sunlight. During the dark they must<br />

use stored food. Plant cells, like animal cells, contain mitochondria in which stored food is converted<br />

to energy by cellular respiration.<br />

Plant cells, which use sunlight for energy and CO 2 for carbon, are said to be autotrophic. In<br />

contrast, animal cells must depend on organic material manufactured by plants for their food. These<br />

are called heterotrophic cells. They act as “middlemen” in the chemical reaction between oxygen<br />

and food material, using the energy from the reaction to carry out their life processes.<br />

SUPPLEMENTARY REFERENCES<br />

Bettelheim, F.A. and March, J., Introduction to Organic and Biochemistry, Saunders College Publishing, Fort<br />

Worth, TX, 1998.<br />

Chesworth, J.M., Stuchbury, T., and Scaife, J.R., An Introduction to Agricultural Biochemistry, Chapman &<br />

Hall, London, 1998.<br />

Garrett, R.H. and Grisham, C.M., Biochemistry, Saunders College Publishing, Philadelphia, 1998.<br />

Gilbert, H.F., Ed., Basic Concepts in Biochemistry, McGraw-Hill, Health Professions Division, New York,<br />

2000.<br />

Kuchel, P.W., Ed., Schaum’s Outline of Theory and Problems of Biochemistry, McGraw-Hill, New York, 1998.<br />

Lea, P.J. and Leegood, R.C., Eds., Plant Biochemistry and Molecular Biology, 2nd ed., John Wiley & Sons,<br />

New York, 1999.<br />

Marks, D.B., Biochemistry, Williams & Wilkins, Baltimore, 1999.<br />

Meisenberg, G. and Simmons, W.H., Principles of Medical Biochemistry, Mosby, St. Louis, 1998.<br />

Switzer, R.L. and Garrity, L.F., Experimental Biochemistry, W.H. Freeman and Co., New York, 1999.<br />

Voet, D., Voet, J.G., and Pratt, C., Fundamentals of Biochemistry, John Wiley & Sons, New York, 1998.<br />

Vrana, K.E., Biochemistry, Lippincott Williams & Wilkins, Philadelphia, 1999.<br />

Wilson, K. and Walker, J.M., Principles and Techniques of Practical Biochemistry, Cambridge University<br />

Press, New York, 1999.<br />

QUESTIONS AND PROBLEMS<br />

1. What is the toxicological importance of lipids? How do lipids relate to hydrophobic (waterdisliking)<br />

pollutants and toxicants?<br />

2. What is the function of a hydrolase enzyme?<br />

3. Match the cell structure on the left with its function on the right:<br />

A. Mitochondria 1. Toxicant metabolism<br />

B. Endoplasmic reticulum 2. Fills the cell<br />

C. Cell membrane 3. DNA<br />

D. Cytoplasm 4. Mediate energy conversion and utilization<br />

E. Cell nucleus 5. Encloses the cell and regulates the passage of materials into and out<br />

of the cell interior<br />

4. The formula of simple sugars is C 6H 12O 6. The simple formula of higher carbohydrates is C 6H 10O 5.<br />

Of course, many of these units are required to make a molecule of starch or cellulose. If higher<br />

carbohydrates are formed by joining together molecules of simple sugars, why is there a difference<br />

in the ratios of C, H, and O atoms in the higher carbohydrates, compared to the simple sugars?<br />

5. Why does wood contain so much cellulose?<br />

6. What would be the chemical formula of a trisaccharide made by the bonding together of three<br />

simple sugar molecules?<br />

33


78 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

7. The general formula of cellulose may be represented as (C 6H 10O 5) x. If the molecular weight of a<br />

molecule of cellulose is 400,000, what is the estimated value of x?<br />

8. During 1 month, a factory for the production of simple sugars, C 6H 12O 6, by the hydrolysis of<br />

cellulose processes 1 million kg of cellulose. The percentage of cellulose that undergoes the<br />

hydrolysis reaction is 40%. How many kilograms of water are consumed in the hydrolysis of<br />

cellulose each month?<br />

9. What is the structure of the largest group of atoms common to all amino acid molecules?<br />

10. Glycine and phenylalanine can join together to form two different dipeptides. What are the<br />

structures of these two dipeptides?<br />

11. One of the ways in which two parallel protein chains are joined together, or cross-linked, is by<br />

way of an –S–S– link. What amino acid to you think might be most likely to be involved in such<br />

a link? Explain your choice.<br />

12. Fungi, which break down wood, straw, and other plant material, have what are called “exoenzymes.”<br />

Fungi have no teeth and cannot break up plant material physically by force. Knowing this, what<br />

do you suppose an exoenzyme is? Explain how you think it might operate in the process by which<br />

fungi break down something as tough as wood.<br />

13. Many fatty acids of lower molecular weight have a bad odor. Speculate as to the reasons why<br />

rancid butter has a bad odor. What chemical compound is produced that has a bad odor? What<br />

sort of chemical reaction is involved in its production?<br />

14. The long-chain alcohol with ten carbons is called decanol. What do you think would be the formula<br />

of decyl stearate? To what class of compounds would it belong?<br />

15. Write an equation for the chemical reaction between sodium hydroxide and cetyl stearate. What<br />

are the products?<br />

16. What are two endocrine glands that are found only in females? Which of these glands is found<br />

only in males?<br />

17. The action of bile salts is a little like that of soap. What function do bile salts perform in the<br />

intestine? Look up the action of soaps, and explain how you think bile salts may function somewhat<br />

like soap.<br />

18. If the structure of an enzyme is illustrated as<br />

how should the structure of its substrate be represented?<br />

19. Look up the structures of ribose and deoxyribose. Explain where the “deoxy” came from in the<br />

name deoxyribose.<br />

20. In what respect are an enzyme and its substrate like two opposite strands of DNA?<br />

21. For what discovery are Watson and Crick noted?<br />

22. Why does an enzyme no longer work if it is denatured?<br />

34


CHAPTER 4<br />

Metabolic Processes<br />

4.1 METABOLISM IN ENVIRONMENTAL BIOCHEMISTRY<br />

The biochemical changes that substances undergo in a living organism are called metabolism.<br />

Metabolism describes the catabolic reactions by which chemical species are broken down by<br />

enzymatic action in an organism to produce energy and components for the synthesis of biomolecules<br />

required for life processes. It also describes the anabolic reactions in which energy is used<br />

to assemble small molecules into larger biomolecules. Metabolism is an essential process for any<br />

organism because it provides the two things essential for life — energy and raw materials.<br />

Metabolism is especially important in toxicological chemistry for two reasons: (1) interference<br />

with metabolism is a major mode of toxic action, and (2) toxic substances are transformed by<br />

metabolic processes to other materials that are usually, though not invariably, less toxic and more<br />

readily eliminated from the organism. This chapter introduces the topic of metabolism in general.<br />

Specific aspects of the metabolism of toxic substances are discussed in Chapter 7.<br />

4.1.1 Metabolism Occurs in Cells<br />

Metabolic processes occur in cells in organisms. Figure 3.1 shows the general structure of<br />

eukaryotic cells in organisms such as animals and fungi. A cell is contained within a cell membrane<br />

composed of a lipid bilayer that separates the contents of the cell from the aqueous medium around<br />

it. Other than the cell nucleus, the material inside the cell is referred to as the cell cytoplasm, the<br />

fluid part of which is the cytosol. The cytosol is an aqueous solution of electrolytes that also contains<br />

enzymes that catalyze some important cell functions, including some metabolic processes. Within<br />

the cytoplasm are specialized organelles that carry out various metabolic functions. Of these,<br />

mitochondria are of particular importance in metabolism because of their role in synthesizing<br />

energetic adenosine triphosphate (ATP) using energy-yielding reactions. Ribosomes are sites of<br />

protein synthesis from mRNA templates (Chapter 8).<br />

4.1.2 Pathways of Substances and Their Metabolites in the Body<br />

In considering metabolic processes, it is important to keep in mind the pathways of nutrients<br />

and xenobiotics in organisms. For humans and other vertebrate animals, materials enter into the<br />

gastrointestinal tract, in which substances are broken down and absorbed into the bloodstream.<br />

Most substances enter the bloodstream through the intestinal walls and are transported first to the<br />

liver, which is the main organ for metabolic processes in the human body. The other raw material<br />

essential for metabolic processes, oxygen from air, enters blood through the lungs. Volatile toxic<br />

substances can enter the bloodstream through the lungs, a major pathway for environmental and<br />

35<br />

79


80 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Liver<br />

Gallbladder<br />

Large intestine<br />

(colon)<br />

Figure 4.1 Major organs involved in digestion.<br />

occupational exposure to xenobiotics. Toxic substances can also be absorbed through the skin.<br />

Undigested food residues and wastes excreted from the liver in bile leave the body through the<br />

intestinal tract as feces. The other major pathway for elimination of waste products from metabolic<br />

processes consists of the kidneys, which remove such materials from blood, and the bladder and<br />

urinary tract through which urine leaves the body. Waste carbon dioxide from the oxidation of food<br />

nutrients is eliminated through the lungs.<br />

4.2 DIGESTION<br />

Stomach<br />

Pancreas<br />

Small intestine<br />

Anus<br />

For most food substances and for a very limited number of toxicants, digestion is necessary<br />

for sorption into the body. Digestion is an enzymatic hydrolysis process by which polymeric<br />

macromolecules are broken down with the addition of water into units that can be absorbed from<br />

the gastrointestinal tract into blood in the circulatory system; material that cannot be absorbed is<br />

excreted as waste, usually after it has been subjected to the action of intestinal bacteria. The digestive<br />

tract and organs associated with it are shown in Figure 4.1. A coating of mucus protects the internal<br />

surface of the digestive tract from the action of the enzymes that operate in it.<br />

Various enzymes perform digestion by acting on materials in the digestive tract. Carbohydrase,<br />

protease, peptidase, lipase, and nuclease enzymes hydrolyze carbohydrates, proteins, peptides,<br />

lipids, and nucleic acids, respectively. Digestion begins in the mouth through the action of amylase<br />

enzyme, which is secreted with saliva and hydrolyzes starch molecules to glucose sugar. The major<br />

enzyme that acts in the stomach is pepsin, a protein-hydrolyzing enzyme secreted into the stomach<br />

as an inactive form (a zymogen) that is activated by a low pH of 1 to 3 in the stomach, resulting<br />

from hydrochloric acid secreted into the stomach. In the small intestine, the digestion of carbohydrates<br />

and proteins is finished, the digestion of fats is initiated, and the absorption of hydrolysis<br />

product nutrients occurs. The first part of the small intestine, the duodenum, is where most digestion<br />

occurs, whereas nutrient absorption occurs in the lower jejunum and ileum. The small intestine<br />

produces a number of enzymes, including aminopeptidase, which converts peptides to other peptides<br />

and amino acids; nuclease; and lactase, which converts lactose (milk sugar) to glactose and glucose.<br />

The liver and the pancreas are not part of the digestive tract as such, but they provide enzymes<br />

and secretions required for digestion to occur in the small intestine. The pancreas secretes amylase,<br />

lipase, and nuclease enzymes, as well as several enzymes involved in breaking down proteins and<br />

peptides. As discussed below with digestion of fats, the liver secretes a substance called bile that<br />

is stored in the gallbladder and then secreted into the duodenum when needed for digestion of fats.<br />

36


METABOLIC PROCESSES 81<br />

By the time that ingested food mass reaches the large intestine or colon, most of the nutrients<br />

have been absorbed. Water and ions are absorbed from the mass of material in the colon, concentrating<br />

it and converting it to a semisolid state. Much of the material in the colon is converted to<br />

bacterial biomass by the action of bacteria, especially Escherichia coli, that metabolize food residues<br />

not digested by humans or animals. These bacteria produce beneficial vitamins, such as Vitamin<br />

K and biotin, that are absorbed through the colon walls and are important in nutrition. The reducing<br />

environment maintained by the bacteria in the colon can reduce some xenobiotics (see the discussion<br />

of metabolic reductions in Section 7.3). One such product is toxic hydrogen sulfide, H 2S, which<br />

is detoxified by special enzymes produced in intestinal wall mucus membranes.<br />

4.2.1 Carbohydrate Digestion<br />

A very simple example of a digestion process is the hydrolysis of sucrose (common table sugar),<br />

H<br />

C<br />

HO<br />

CH 2 OH<br />

C O<br />

H<br />

OH H<br />

C C<br />

H OH<br />

H<br />

C<br />

HO<br />

O<br />

HOCH2 H<br />

C C<br />

O H HO<br />

C C<br />

H<br />

H2O C<br />

CH2OH HO H<br />

Sucrose<br />

CH2OH C O<br />

H<br />

H<br />

C<br />

HOCH2 C<br />

O<br />

OH<br />

C<br />

OH<br />

C<br />

H<br />

C<br />

OH H<br />

H<br />

C<br />

HO<br />

C<br />

CH2OH H OH<br />

HO H<br />

Glucose Fructose<br />

(4.2.1)<br />

to produce glucose and fructose monosaccharides that can be absorbed through intestine walls to<br />

undergo metabolism in the body. Each digestive hydrolysis reaction of carbohydrates has its own<br />

enzyme. Sucrase enzyme carries out the reaction above, whereas amylase enzyme converts starch<br />

to a disaccharide with two glucose molecules called maltose, and maltose in turn is hydrolyzed to<br />

glucose by the action of maltase enzyme. A third important disaccharide is lactose or “milk sugar,”<br />

each molecule of which is hydrolyzed by digestive processes to give a molecule of glucose and<br />

one of galactose.<br />

HO<br />

C<br />

H<br />

CH2OH C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

OH<br />

C<br />

H<br />

H OH<br />

37<br />

Galactose


82 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

O H<br />

CH3 (CH2 ) 16C O C H<br />

O<br />

O H C C C(CH2 ) 16CH3 CH3 (CH2 ) 16C O C H<br />

Triglyceride<br />

(fat)<br />

Lipase<br />

enzyme<br />

H<br />

O H<br />

CH3 (CH2 ) 16C O C H<br />

Fatty<br />

acids<br />

Figure 4.2 Illustration of digestion of fats (triglycerides).<br />

Digestion can be a limiting factor in the ability of organisms to utilize saccharides. Many adults<br />

lack the lactase enzyme required to hydrolyze lactose. When these individuals consume milk<br />

products, the lactose remains undigested in the intestine, where it is acted upon by bacteria. These<br />

bacteria produce gas and intestinal pain, and diarrhea may result. The lack of a digestive enzyme<br />

for cellulose in humans and virtually all other animals means that these animals cannot metabolize<br />

cellulose. The cellulosic plant material eaten by ruminant animals such as cattle is actually digested<br />

by the action of enzymes produced by specialized rumen bacteria in the stomachs of such animals.<br />

4.2.2 Digestion of Fats<br />

Fats and oils are the most common lipids that are digested. Digestion breaks fats down from<br />

triglycerides to di- and monoglycerides, fatty acids and their salts (soaps) and glycerol, which pass<br />

through the intestine wall, where they are resynthesized to triglycerides and transported to the blood<br />

through the lymphatic system (see Figure 4.2).<br />

A special consideration in the digestion of fats is that they are not water soluble and cannot be<br />

placed in aqueous solution along with the water-soluble lipase digestive enzymes. However, intimate<br />

contact is obtained by emulsification of fats through the action of bile salts from glycocholic and<br />

taurocholic acids produced from cholesterol in the liver:<br />

HO<br />

H C OH<br />

H C OH H C<br />

O<br />

O<br />

CH3 (CH2 ) 16C O<br />

Diglyceride<br />

C<br />

H<br />

H CH3 (CH2 ) 16C Monoglyceride<br />

O C<br />

H<br />

H HO<br />

Glycerol<br />

C<br />

H<br />

+ + +<br />

{<br />

O<br />

CH3 (CH2 ) 16C OH<br />

O<br />

CH3 (CH2 ) 16C Reassembly of fat<br />

digestion products<br />

38<br />

HO<br />

OH<br />

O<br />

C<br />

H<br />

C<br />

H<br />

O<br />

CH3 (CH2 ) 16C O - Na +<br />

H<br />

C<br />

Triglycerides<br />

H<br />

OH<br />

H<br />

OH<br />

Representation of a bile salt<br />

showing steroid skeleton<br />

from cholesterol<br />

}


METABOLIC PROCESSES 83<br />

H<br />

H<br />

N +<br />

H<br />

H<br />

H<br />

C<br />

CH 3<br />

N + H H O<br />

H<br />

C C<br />

CH 3<br />

Figure 4.3 Illustration of the enzymatic hydrolysis of a tetrapeptide such as occurs in the digestion of protein.<br />

4.2.3 Digestion of Proteins<br />

O H H O H H O H H O<br />

C N C C N C C N C C<br />

H H C H H C H<br />

H C<br />

S<br />

H<br />

O -<br />

Digestion of proteins occurs by enzymatic hydrolysis in the small intestine (Figure 4.3). The<br />

digestion of protein produces single amino acids. These can enter the bloodstream through the<br />

small intestine walls. The amino acids circulate in the bloodstream until further metabolized or<br />

used for protein synthesis; there is not a “storage depot” for amino acids as there is for lipids,<br />

which are stored in “fat depots” in adipose tissue. However, the body does break down protein<br />

tissue (muscle) to provide amino acids in the bloodstream.<br />

4.3 METABOLISM OF CARBOHYDRATES, FATS, AND PROTEINS<br />

In the preceding section the digestion of carbohydrates, fats, and proteins by the enzymatic<br />

hydrolysis of their molecules was discussed. Digestion enables these materials to enter the bloodstream<br />

as relatively small molecules. Once in the bloodstream, these small molecules undergo<br />

further metabolic reactions to enable their use for energy production and tissue synthesis. These<br />

metabolic processes are all rather complex and beyond the scope of this chapter. However, the main<br />

points are covered below.<br />

4.3.1 An Overview of Catabolism<br />

+ 3H 2 O<br />

The overall process by which energy-yielding nutrients are broken down to provide the energy<br />

required for muscle movement, protein synthesis, nerve function, maintenance of body heat, and<br />

other energy-consuming functions is illustrated in Figure 4.4. The approximate empirical formula<br />

of the biomolecules from which energy is obtained in catabolism can be represented as {CH 2O}.<br />

The overall energy-yielding catabolic process is the following:<br />

{CH 2O} + O 2 → CO 2 + H 2O + energy (4.3.1)<br />

Figure 4.4 as summarized in Reaction 4.3.1 represents oxidative respiration, in which glucose,<br />

other nutrients that can be converted to glucose, and the intermediates that glucose generates are<br />

oxidized completely to carbon dioxide and water, yielding large amounts of energy. Oxidative<br />

CH 3<br />

O -<br />

+ N C C + +<br />

+<br />

H O -<br />

H H O<br />

N C C<br />

H H C H<br />

+<br />

H O -<br />

H H O<br />

H<br />

H<br />

39<br />

N + H H O<br />

C C<br />

H<br />

H C H<br />

H C<br />

S<br />

H<br />

CH 3<br />

O -


84 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Glycolysis: degradation and partial oxidation<br />

Fatty acids and<br />

glycerol from digestion<br />

of triglycerides<br />

Fatty acids<br />

Oxidation of nutrients to CO 2 and H 2 O<br />

Electron transport<br />

chain<br />

ATP<br />

Glycerol<br />

CO 2<br />

Glucose, fructose, and<br />

galactose from digestion<br />

of polysaccarides<br />

Glucose ATP<br />

Pyruvate<br />

Amino acids<br />

from digestion<br />

of proteins<br />

Acetyl = CoA Transamination<br />

Citric acid<br />

cycle<br />

NADH FADH 2<br />

Figure 4.4 Overview of catabolic metabolism, the process by which nutrients are broken down to provide<br />

energy.<br />

respiration is in fact a very complicated process involving many steps, numerous enzymes, and a<br />

variety of intermediate species. Discussed in more detail in Section 4.4, oxidative respiration in<br />

eukaryotic organisms begins with the conversion of glucose to pyruvic acid, a step that does not<br />

require oxygen. The second stage of oxidative respiration is the conversion of pyruvic acid to acetyl<br />

coenzyme A (acetyl-CoA). In the third stage, the acetyl-CoA goes through the citric acid cycle, in<br />

which chemical bond energy harvested in the oxidation of the biomolecules metabolized is converted<br />

primarily to a species designated as NADH. In the last stage of oxidative respiration, NADH<br />

ATP<br />

O 2 H 2 O ADP<br />

40<br />

NH 4 + , urea


METABOLIC PROCESSES 85<br />

transfers electrons to molecular O 2 and generates high-energy species (ATP) that are utilized for<br />

metabolic needs.<br />

4.3.2 Carbohydrate Metabolism<br />

As discussed in the preceding section, starch and the major disaccharides are broken down by<br />

digestive processes to glucose, fructose, and galactose monosaccharrides. Fructose and galactose<br />

are readily converted by enzyme action to glucose. Glucose is converted to the glucose 1-phosphate<br />

species:<br />

H<br />

C<br />

HO<br />

From the glucose 1-phosphate form, glucose may be incorporated into macromolecular (polymeric)<br />

glycogen for storage in the animal’s body and to provide energy-producing glucose on<br />

demand. For the production of energy, the glucose 1-phosphate enters the catabolic process through<br />

glycolysis, discussed in Section 4.4.<br />

4.3.3 Metabolism of Fats<br />

Fats are stored and circulated through the body as triglycerides, which must undergo hydrolysis<br />

to glycerol and fatty acids before they are further metabolized. Glycerol is broken down via the<br />

glycolysis pathway discussed above for carbohydrate metabolism. The fatty acids are broken down<br />

in the fatty acid cycle, in which a long-chain fatty acid goes through a number of sequential steps<br />

to be shortened by two carbon fragments, producing CO 2, H 2O, and energy.<br />

4.3.4 Metabolism of Proteins<br />

O -<br />

A central feature of protein metabolism is the amino acid pool, consisting of amino acids in<br />

the bloodstream. Figure 4.5 illustrates the metabolic relationship of the amino acid pool to protein<br />

breakdown, synthesis, and storage.<br />

Proteins are synthesized from amino acids in the amino acid pool as discussed in Section 3.3.<br />

This occurs through the joining of H3N + –<br />

– and –CO2 groups at peptide bonds, with the elimination<br />

of H2O for each peptide bond formed. The body can make many of the amino acids it needs, but<br />

eight of them, the essential amino acids, cannot be synthesized in the human body and must be<br />

included in the diet.<br />

The first step in the metabolic breakdown of amino acids is often the replacement of the – NH2 group with a C=O group by the action of α-ketoglutaric acid in a process called transamination.<br />

Oxidative deamination then regenerates the α-ketoglutaric acid from the glutamic acid product<br />

of transamination. These processes are illustrated in Figure 4.6. As a net result of transamination,<br />

N(-III) is removed from amino acids and eliminated from the body. For this to occur, nitrogen is<br />

first converted to urea:<br />

O -<br />

O<br />

CH2OH C O<br />

P<br />

O<br />

H<br />

OH<br />

C<br />

H<br />

C<br />

C Glucose 1-phosphate<br />

OH<br />

H OH<br />

41


86 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

Fats<br />

NH +<br />

4<br />

Carbohydrates<br />

Metabolic products,<br />

O<br />

H2N–C–NH2 ,<br />

CO2 , H2O, energy<br />

=<br />

Digested proteins<br />

Amino<br />

acids<br />

Amino acid pool<br />

Figure 4.5 Main features of protein metabolism.<br />

O<br />

H2N C<br />

C H<br />

CH 3<br />

Amino acid<br />

(alanine)<br />

OH<br />

O<br />

C OH<br />

+ O C<br />

H C H<br />

H C H<br />

HO C<br />

O<br />

α-ketoglutaric acid<br />

+<br />

NH 4<br />

Amino acids<br />

Nitrogenous compounds other<br />

than protein, such as heme in<br />

blood hemoglobin, nitrogenous<br />

bases in nucleic acids,<br />

and creatinine<br />

To citric acid cycle<br />

O<br />

O<br />

C OH C OH<br />

O C + H2N C H<br />

CH3 H C H<br />

Pyruvic acid,<br />

α-keto acid<br />

H<br />

HO<br />

C<br />

C<br />

H<br />

Glutamic acid<br />

O<br />

+ + {O}<br />

Body protein<br />

(muscle tissue,<br />

enzymes)<br />

Figure 4.6 Transamination of an amino acid and regeneration of α-ketoglutaric acid by oxidative deamination.<br />

Urea is a solute that is contained in urine, and it is eliminated from the body via the kidneys and<br />

bladder.<br />

The α-keto acids formed by transamination of amino acids are further broken down in the citric<br />

acid (Krebs) cycle. This process yields energy, and the body’s energy needs can be met with protein<br />

if sufficient carbohydrates or fats are not available.<br />

H +<br />

H<br />

O<br />

H<br />

N C N<br />

H H<br />

42<br />

Urea


METABOLIC PROCESSES 87<br />

4.4 ENERGY UTILIZATION BY METABOLIC PROCESSES<br />

Energy in the form of free energy needed by organisms is provided by enzymatically mediated<br />

oxidation–reduction reactions. Oxidation in a biological system, as in any chemical system, is the<br />

loss of electrons, and reduction is the gain of electrons. A species that is oxidized by losing a<br />

negatively charged electron may maintain electrical neutrality by losing H + ion; the loss of both e –<br />

and H + is equivalent to the loss of a hydrogen atom, H.<br />

A large number of steps within several major cycles are involved in energy conversion, transport,<br />

and utilization in organisms. It is beyond the scope of this book to discuss all of these mechanisms<br />

in detail. However, it is useful to be aware of the main mechanisms involving energy in relation<br />

to biochemical processes in which chemical or photochemical energy is utilized by organisms.<br />

They are the following:<br />

• Glycolysis, in which, through a series of enzymatic reactions, a six-carbon glucose molecule is<br />

converted to two three-carbon pyruvic acid (pyruvate) species with the release of a relatively small<br />

amount of the energy in the glucose<br />

• Cellular respiration, which occurs in the presence of molecular oxygen, O 2, and involves the<br />

conversion of pyruvate to carbon dioxide, CO 2, with the release of relatively large amounts of<br />

energy by way of intermediate chemical species<br />

• Fermentation, which occurs in the absence of molecular O 2 and produces energy-rich molecules,<br />

such as ethanol or lactic acid, with release of relatively little useable energy<br />

4.4.1 High-Energy Chemical Species<br />

Metabolic energy is provided by the breakdown and oxidation of energy-providing nutrients,<br />

especially glucose. Usually, however, the energy is needed in a different location and at a different<br />

time from the place and time where it is generated. This entails the synthesis of high-energy<br />

chemical species that require energy for their synthesis and release it when they break down. Of<br />

these, the most important is ATP:<br />

- O<br />

O O O<br />

P O P O P<br />

O -<br />

O -<br />

O -<br />

which is generated by the addition of inorganic phosphate, commonly represented as P i, from<br />

adenosine diphosphate (ADP):<br />

- O<br />

O O<br />

P O P<br />

O -<br />

O -<br />

N<br />

NH 2<br />

H O<br />

N<br />

O C<br />

H C<br />

H<br />

H<br />

C<br />

H<br />

C<br />

C<br />

H<br />

HO OH<br />

H<br />

O C<br />

H C<br />

H<br />

O<br />

H<br />

C<br />

H<br />

C<br />

N<br />

C<br />

H<br />

HO OH<br />

43<br />

N<br />

NH 2<br />

N<br />

N<br />

ADP<br />

N<br />

N<br />

ATP<br />

C<br />

C<br />

H<br />

H


88 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

- O<br />

- O<br />

O<br />

P<br />

O<br />

P<br />

O<br />

O<br />

O<br />

Figure 4.7 Structural formula of nicotinamide adenine dinucleotide in its reduced form of NADH + H + and its<br />

oxidized form NAD + .<br />

When ATP releases inorganic phosphate and reverts to ADP, a quantity of energy equivalent to 31<br />

kJ of energy per mole of ATP is released that can be utilized metabolically. A pair of species that<br />

are similar in function to ATP and ADP are guanine triphosphate (GTP) and guanine diphosphate<br />

(GDP).<br />

An important aspect of enzymatic oxidation–reduction reactions involves the transfer of hydrogen<br />

atoms. This transfer is mediated by coenzymes (substances that act together with enzymes)<br />

nicotinamide adenine dinucleotide (NAD) and nicotinamide adenine dinucleotide phosphate<br />

(NADP). These two species pick up H atoms to produce NADH and NADPH, respectively, both<br />

of which can function as hydrogen atom donors. Another pair of species involved in oxidation–reduction<br />

processes by hydrogen atom transfer consists of flavin adenine triphosphate (FAD)<br />

and its hydrogenated form FADH 2. The structural formulas of NAD and its cationic form, NAD + ,<br />

are shown in Figure 4.7.<br />

4.4.2 Glycolysis<br />

H<br />

CH<br />

HO<br />

O<br />

H H<br />

H H<br />

H<br />

C H O<br />

H<br />

H<br />

Reduced form NADH + H<br />

H H<br />

Oxidation<br />

H<br />

+ Reduced form NAD +<br />

OH<br />

H<br />

N<br />

HO OH<br />

H<br />

N<br />

O<br />

C<br />

H +<br />

N N<br />

H<br />

N<br />

H<br />

NH 2<br />

Reduction<br />

+ 2H<br />

Nicotinamide adenine dinucleotide<br />

Glycolysis is a multistepped, anaerobic (without oxygen) process in which a molecule of glucose<br />

is broken down in the absence of O 2 to produce two molecules of pyruvic acid (pyruvate anion)<br />

and energy. Glycolysis occurs in cell protoplasm and may be followed by either cellular respiration<br />

utilizing O 2 or fermentation in the absence of O 2. The glycolysis of a molecule of glucose results<br />

in the net formation of two molecules of energetic ATP and the reduction of two NAD + to two<br />

molecules of NADH plus H + .<br />

The first part of the glycolysis process consumes energy provided by the conversion of two<br />

ATPs to ADP. It consists of five major steps in which a glucose molecule is converted to two<br />

glylceraldehyde 3-phosphate molecules with intermediate formation of glucose 6-phosphate, fructose<br />

6-phosphate, fructose 1,6-biphosphate, and dihydroxyacetone:<br />

44<br />

N<br />

+<br />

O<br />

C<br />

H<br />

N<br />

H


METABOLIC PROCESSES 89<br />

CH2OH H<br />

C<br />

HO<br />

C<br />

H<br />

OH<br />

C<br />

O<br />

H<br />

C<br />

2ATP 2ADP + Pi H<br />

C<br />

Five steps, energy consumed<br />

OH<br />

H OH<br />

Glucose<br />

(4.4.1)<br />

The second part of the glycolyis process is the five-step conversion of glyceraldehyde to pyruvate<br />

accompanied by conversion of four ADPs to four ATPs:<br />

H<br />

O<br />

H C<br />

2<br />

HO C H<br />

C O<br />

H<br />

(4.4.2)<br />

Since two molecules of ATP are converted to ADP in the first part of the glycolysis process,<br />

there is a net gain of two molecules of ATP. The second part of the glycolysis process also yields<br />

two molecules of NADH + H + per molecule of glucose. Subsequently, the energy-yielding conversion<br />

of the two molecules of ATP back to ADP and the oxidation of NADH,<br />

2NADH + 2H + + O 2 → NAD + + H 2O + energy (4.4.3)<br />

can provide energy for metabolic needs. The three conversions accomplished in glycolysis are (1)<br />

glucose to pyruvate, (2) ADP to ATP, and (3) NAD + to NADH. The net reaction for glycolysis may<br />

be summarized as<br />

Glucose + 2ADP + 2P i + 2NAD + → 2Pyruvate + 2ATP + 2NADH + 2H + + 2H 2O (4.4.4)<br />

In addition to glucose, other monosaccharides and nutrients may be converted to intermediates in<br />

the glycolysis cycle and enter the cycle as these intermediates.<br />

4.4.3 Citric Acid Cycle<br />

P O -<br />

O<br />

O<br />

Glyceraldehyde<br />

3-phosphate<br />

4ADP + P i 4ATP<br />

Five steps, energy released<br />

2NAD + 2NADH + 2H +<br />

P O -<br />

H<br />

O<br />

H<br />

2<br />

HO<br />

C<br />

O O<br />

C H<br />

C O<br />

H<br />

Glyceraldehyde<br />

3-phosphate<br />

Glycolysis yields a relatively small amount of energy. Much larger amounts of energy may be<br />

obtained by complete oxidation of bionutrients to CO 2 and 2H 2O, which occurs in heterotrophic<br />

organisms that utilize oxygen for respiration. The pyruvic acid product of glycolysis can be oxidized<br />

to the acetyl group, which becomes bound to coenzyme A in the highly energetic molecule acetyl-<br />

CoA, as shown by the following reaction:<br />

45<br />

2<br />

O<br />

O<br />

C<br />

C<br />

C<br />

H<br />

O -<br />

H H<br />

Pyruvate


90 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

O<br />

C<br />

C<br />

OH<br />

O<br />

H C<br />

H<br />

H<br />

S CoA<br />

+ HS-CoA C O + CO2 + 2{H}<br />

H C<br />

H<br />

H<br />

(4.4.5)<br />

Acetyl-CoA enters the citric acid cycle (also called the Krebs cycle), which occurs in cell mitochondria.<br />

In the Krebs cycle, the acetyl group is oxidized to CO 2 and water, harvesting a substantial<br />

amount of energy. This complex cycle starts with the reaction of oxaloacetate with acetyl-CoA,<br />

O C O<br />

Oxaloacetate Acetyl = CoA Citrate<br />

-<br />

O C<br />

C<br />

C O -<br />

H<br />

O<br />

H<br />

+<br />

H<br />

S CoA<br />

C O<br />

C H<br />

H<br />

C<br />

C<br />

C O -<br />

HO<br />

H<br />

O<br />

C O<br />

H<br />

-<br />

O C O<br />

H C H<br />

O<br />

-<br />

(4.4.6)<br />

to produce citrate. In a series of steps involving a number of intermediates, CO 2 is evolved and H<br />

is removed by NAD + to yield NADH + H + and by FAD to yield FADH 2. Guanosine triphosphate<br />

is also generated from guanosine diphosphate in the citric acid cycle and later generates ATP. The<br />

anions of several four-, five-, and six-carbon organic acids are generated as intermediates in the<br />

citric acid cycle, including citric, isocitric, ketoglutaric, succinic, fumaric, malic, and oxaloacetic<br />

acids; the last of these reacts with additional acetyl-CoA from glycolysis to initiate the cycle again.<br />

Structural formulas of the intermediate species generated in the citric acid cycle are shown in<br />

Figure 4.8. The reduced carrier molecules generated in the citric acid cycle, NADH and FADH 2,<br />

are later oxidized in the respiratory chain (see below) to produce ATP and are, therefore, the major<br />

conduits of energy from the citric acid cycle. The reaction for one complete cycle of the citric acid<br />

cycle can be summarized as follows:<br />

Acetyl-CoA + 3NAD + + FAD + GDP + P i + 2H 2O →<br />

3NADH + FADH 2 + 2CO 2 + 2H + + GTP + HS-CoA (4.4.7)<br />

4.4.4 Electron Transfer in the Electron Transfer Chain<br />

As indicated by Reaction 4.3.1, the driving force behind the high-energy yields of oxidative<br />

respiration is the reaction with molecular oxygen to produce H 2O. Electrons removed from glucose<br />

and its products during oxidative respiration are donated to O 2, which is the final electron receptor.<br />

To this point in the discussion of oxidative respiration, molecular oxygen has not entered any of<br />

the steps. It does so during transfer of electrons in the electron transfer chain, the step at which<br />

most of the energy is harvested from oxidative respiration. Electrons picked up from glycolysis<br />

and citric acid cycle intermediates are transferred to the electron transport chain via NAD + /(NADH<br />

+ H + ) and FAD/FADH 2. The overall process, which occurs in several small increments, is represented<br />

for NADH as<br />

NADH + H + + ½O 2 → NAD + + H 2O (4.4.8)<br />

46


METABOLIC PROCESSES 91<br />

H<br />

H C<br />

H<br />

O<br />

O<br />

C CoA HO C C O<br />

Acetyl = CoA<br />

-<br />

O<br />

H<br />

C<br />

C<br />

O<br />

H<br />

-<br />

H C H<br />

C O -<br />

O<br />

C C O<br />

O<br />

-<br />

O<br />

H<br />

C<br />

C<br />

O<br />

H<br />

-<br />

C H<br />

C O -<br />

H<br />

C<br />

HO<br />

C<br />

C O<br />

O<br />

-<br />

O<br />

H H<br />

O<br />

O<br />

-<br />

From glycolysis<br />

O<br />

C<br />

H C H<br />

Citrate Isocitrate α-Ketoglutarate<br />

O<br />

H C H<br />

O C<br />

S<br />

-<br />

O<br />

O C<br />

H C H<br />

H C H<br />

CoA<br />

-<br />

O<br />

C<br />

H C H<br />

C O<br />

O<br />

-<br />

O<br />

C<br />

-<br />

O<br />

C<br />

C<br />

C O<br />

O<br />

-<br />

O<br />

H<br />

C<br />

H<br />

-<br />

O<br />

C<br />

C<br />

C O<br />

O<br />

-<br />

O<br />

H OH<br />

C H H<br />

-<br />

O<br />

C<br />

C<br />

C O<br />

O<br />

-<br />

O<br />

H H<br />

Oxaloacetate Malate Fumarate Succinate Succinyl CoA<br />

Figure 4.8 Intermediates in the citric acid cycle shown in the ionized forms in which they exist at physiological<br />

pH values. The final oxaloacetate product reacts with acetyl-CoA from glycolysis to start the cycle<br />

over again.<br />

The electron transfer chain converts ADP to highly energetic ATP, which provides energy in cells.<br />

By occurring in several reactions, the oxidation of NADH + H + to NAD + releases energy in small<br />

increments enabling its efficient utilization.<br />

4.4.5 Electron Carriers<br />

Electron carriers are chemical species that exist in both oxidized and reduced forms capable<br />

of reversible exchange of electrons. Electron carriers consist of flavins, coenzyme Q, iron–sulfur<br />

proteins, and cytochromes. As shown in Figure 4.9, cytochromes contain iron bound with four N<br />

atoms attached to protein molecules, a group called the heme group. The iron ions in cytochromes<br />

are capable of gaining and losing electrons to produce Fe 2+ and Fe 3+ , respectively. Interference with<br />

the action of cytochromes is an important mode of the action of some toxicants. Cyanide ion, CN – ,<br />

has a strong affinity for Fe 3+ in ferricytochrome, preventing it from reverting back to the Fe 2+ form,<br />

thus stopping the transfer of electrons to O 2 and resulting in rapid death in the case of cyanide<br />

poisoning.<br />

4.4.6 Overall Reaction for Aerobic Respiration<br />

From the discussion above, it is obvious that aerobic respiration is a complex process involving<br />

a multitude of steps and a large number of intermediate species. These can be summarized by the<br />

following overall net reaction for the catabolic metabolism of glucose:<br />

C 6H 12O 6 + 10NAD + + 2FAD + + 36ADP + 36P i + 14H + + 6O 2 →<br />

6CO 2 + 36ATP + 6H 2O + 10NADH + 6FADH 2<br />

47<br />

(4.4.9)


92 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

4.4.7 Fermentation<br />

HC<br />

HC<br />

Figure 4.9 Heme group in a cytochrome involved in electron exchange.<br />

N<br />

N<br />

Fe 3+<br />

N<br />

H<br />

C<br />

N<br />

Fermentation occurs when O 2 is not utilized for aerobic respiration through the citric acid<br />

cycle and the electron transport chain. Glycolysis still occurs as a prelude to fermentation with<br />

some production of ATP, but the utilization of energy from glucose is much less than when aerobic<br />

respiration occurs. Instead of complete oxidation of glucose to carbon dioxide and water, fermentation<br />

stops with an organic molecule. Fermentation is carried out by a variety of bacteria and by<br />

eukaryotic cells in the human body. Muscle cells carry out fermentation of pyruvate to lactate under<br />

conditions of insufficient oxygen, and the accumulation of lactate in muscle cells is responsible<br />

for the pain associated with extreme exertion. Nerve cells are incapable of carrying out fermentation,<br />

which is why brain tissue is rapidly destroyed when the brain is deprived of oxygen.<br />

Lactic acid fermentation occurs when lactate is the end product of fermentation. Coupled<br />

with glycolysis, lactic acid fermentation can generate ATP from ADP and provide energy for cellular<br />

processes. The fermentation step in lactic acid fermentation generates NAD + from NADH + H + ,<br />

and the NAD + cycles back to the glycolysis process. The lactic acid fermentation cycle is illustrated<br />

in Figure 4.10.<br />

Alcoholic fermentation occurs when the end product is ethanol, as shown in Figure 4.11. In<br />

this process the pyruvate is first converted enzymatically to acetaldehyde. The conversion of<br />

acetaldehyde to ethanol produces NAD + from NADH + H + , and the NAD + is cycled through the<br />

glycolysis process. As with lactic acid fermentation, the glycolysis process produces usable energy<br />

contained in two molecules of ATP produced for each molecule of glucose metabolized.<br />

4.5 USING ENERGY TO PUT MOLECULES TOGETHER: ANABOLIC REACTIONS<br />

The preceding section has discussed in some detail the complicated processes by which complex<br />

molecules are disassembled to extract energy for metabolic needs. Much of this energy goes into<br />

anabolic metabolic processes to put small molecules together to produce large molecules needed<br />

for function and structure in organisms. Typical of the small molecules so put together are glucose<br />

monosaccharide molecules, assembled into starch macromolecules, and amino acids, assembled<br />

into proteins. In all cases of macromolecule synthesis, an H atom is removed from one molecule<br />

and an –OH group from the other to link the two together:<br />

S<br />

48<br />

CH<br />

CH<br />

Fe 2+<br />

H<br />

C<br />

Protein<br />

S


METABOLIC PROCESSES 93<br />

Fermentation Glycolysis<br />

H<br />

C<br />

HO<br />

CH2OH C O 2ADP 2ATP<br />

H<br />

H<br />

C<br />

OH H<br />

2<br />

OH<br />

C C<br />

Pyruvate<br />

H OH 2NAD<br />

Glucose<br />

+ 2NADH + 2H +<br />

O<br />

O C<br />

O C<br />

H C H<br />

H<br />

2<br />

-<br />

O O<br />

C<br />

C<br />

C<br />

-<br />

H OH<br />

Lactate<br />

H<br />

H<br />

H<br />

Figure 4.10 Lactic acid fermentation in which the conversion of pyruvate to lactate is coupled with glycolysis<br />

to produce energetic ATP.<br />

Fermentation Glycolysis<br />

CH2OH C O 2ADP 2ATP<br />

H<br />

H<br />

H<br />

C C<br />

OH H<br />

2<br />

HO<br />

OH<br />

C C<br />

Pyruvate<br />

H OH 2NAD<br />

Glucose<br />

+ 2NADH + 2H +<br />

O<br />

O C<br />

O C<br />

H C H<br />

H<br />

-<br />

2CO 2<br />

H H<br />

O H<br />

2 HO C C H<br />

2 C C H<br />

H H<br />

H<br />

H<br />

Ethanol Acetaldehyde<br />

Figure 4.11 Alcoholic fermentation in which the conversion of pyruvate to ethanol through an acetaldehyde<br />

intermediate is coupled with glycolysis to produce energetic ATP.<br />

{Molecule A}–H + HO–{Molecule B} → {Molecule A}–{Molecule B} + H 2O (4.5.1)<br />

Since a molecule of water is eliminated for each linkage formed, the anabolic reactions leading to<br />

macromolecule formation are dehydration reactions. Recall from Section 4.2 that when macromolecular<br />

nutrient molecules are digested prior to their entering the body’s system, a molecule of<br />

water is added for each linkage broken — a hydrolysis reaction. The energy required for the<br />

anabolic synthesis of macromolecules is provided by catabolic processes of glycolysis, the citric<br />

acid cycle, and electron transport.<br />

A variety of macromolecules are produced anabolically. The more important of these are listed<br />

below:<br />

• Polysaccharide glycogen (animals) and starch (plants) produced for energy storage from glucose<br />

49


94 TOXICOLOGICAL CHEMISTRY AND BIOCHEMISTRY<br />

• Polysaccharide chitin composing shells of crabs and similar creatures produced from modified<br />

glucose<br />

• Lipid triglyceride fats and oils used for energy storage produced from glycerol and three fatty acids<br />

• Phospholipids present in cell membranes produced from glycerol, two fatty acids, and phosphate<br />

• Globular proteins (in enzymes) and structural proteins (in muscle) produced from amino acids<br />

• Nucleic acids that provide the genetic code and directions for protein synthesis composed of<br />

phosphate, nitrogenous bases, deoxyribose (DNA), and oxyribose (RNA)<br />

The anabolic processes by which macromolecules are produced are obviously important in life<br />

processes. Remarkably, these processes generally occur properly, making the needed materials<br />

when and where needed. However, in some cases things go wrong with potentially catastrophic<br />

results. This can occur through the action of toxicants and is a major mode of the action of toxic<br />

substances.<br />

4.6 METABOLISM AND TOXICITY<br />

Metabolism is of utmost importance in toxicity. Details of the metabolism of toxic substances<br />

and their precursors are addressed in Chapter 7, “Toxicological Chemistry.” At this point it should<br />

be noted that there are several major aspects of the relationship between toxic substances and<br />

metabolism, as listed below:<br />

• Some substances that are not themselves toxic are metabolized to toxic species. Most substances<br />

regarded as causing cancer must be metabolically activated to produce species that are the ultimate<br />

carcinogenic agents.<br />

• Toxic species are detoxified by metabolic processes.<br />

• Metabolic processes act to counter the effects of toxic substances.<br />

• Metabolic processes of fungi, bacteria, and protozoa act to degrade toxic substances in the water<br />

and soil environments.<br />

• Adverse effects on metabolic processes constitute a major mode of action of toxic substances. For<br />

example, cyanide ion bonds with ferricytochrome oxidase, a form of an enzyme containing iron(III)<br />

that cycles with ferrouscytochrome oxidase, containing iron(II), in the respiration process by which<br />

molecular oxygen is utilized, thus preventing the utilization of O 2 and leading to rapid death.<br />

4.6.1 Stereochemistry and Xenobiotics Metabolism<br />

Recall from Section 1.9 that some molecules can exist as chiral enantiomers that are mirror<br />

images of each other. Although enantiomers may appear to be superficially identical, they may<br />

differ markedly in their metabolism and toxic effects. Much of what is known about this aspect of<br />

xenobiotics has been learned from studies of the metabolism and effects of pharmaceuticals. For<br />

example, one of the two enantiomers that comprise antiepileptic Mesantoin is much more rapidly<br />

hydroxylated in the body and eliminated than is the other enantiomer. The human cytochrome P-<br />

450 enzyme denoted CYP2D6 is strongly inhibited by quinidine, but is little affected by quinine,<br />

an optical isomer of quinidine. Cases are known in which a chiral secondary alcohol is oxidized<br />

to an achiral ketone, and then reduced back to the secondary alcohol in the opposite configuration<br />

of the initial alcohol.<br />

50


METABOLIC PROCESSES 95<br />

SUPPLEMENTARY REFERENCES<br />

Brody, T., Nutritional Biochemistry, 2nd. ed., Academic Press, San Diego, 1999.<br />

Finley, J.W. and Schwass, D.E., Eds., Xenobiotic Metabolism, American Chemical Society, Washington, D.C.,<br />

1985.<br />

Groff, J.L., Gropper, S.S., and Hunt, S.M., Advanced Nutrition and Human Metabolism, 2nd ed., West<br />

Publishers, Minneapolis/St. Paul, 1995.<br />

Hutson, D.H., Caldwell, J., and Paulson, G.D., Intermediary Xenobiotic Metabolism in Animals: Methodology,<br />

Mechanisms, and Significance, Taylor & Francis, London, 1989.<br />

Illing, H.P.A., Ed., Xenobiotic Metabolism and Disposition, CRC Press, Boca Raton, FL, 1989.<br />

Salway, J.G., Metabolism at a Glance, 2nd ed., Blackwell Science, Malden, MA, 1999.<br />

Stephanopoulos, G., Aristidou, A.A., and Nielsen, J., Metabolic Engineering Principles and Methodologies,<br />

Academic Press, San Diego, 1998.<br />

QUESTIONS AND PROBLEMS<br />

1. Define metabolism and its relationship to toxic substances.<br />

2. Distinguish between digestion and metabolism. Why is digestion relatively unimportant in regard<br />

to toxic substances, most of which are relatively small molecules?<br />

3. What is the fundamental difference between the digestion of fats and that of complex carbohydrates<br />

(starches) and proteins? What role is played by bile salts in the digestion of fats?<br />

4. What are the functions of ADP, ATP, NAD + , and NADH in metabolism?<br />

5. What is the overall reaction mediated by the Krebs cycle? What does it produce that the body needs?<br />

6. What is the amino acid pool? What purposes does it serve?<br />

7. What is meant by an essential amino acid?<br />

8. What is transamination? What product of amino acid synthesis is eliminated from the body by the<br />

kidneys?<br />

9. Give the definition and function of an energy carrier species in metabolism.<br />

51


BYLAAG 2<br />

Clark, JH and Macquarrie, D, 2002, Handbook of Green Chemistry And Technology, 2 nd ed., Blackwell<br />

Science: Oxford, p. 1 – 28.<br />

52


1 Introduction<br />

1.1 Chemistry—past, present and future<br />

Chapter 1: Introduction<br />

Chemistry is having a difficult time. While society<br />

continues to demand larger quantities of increasingly<br />

sophisticated chemical products, it also regards the<br />

industries that manufacture these products with<br />

increasing degrees of suspicion and fear.<br />

The range of chemical products in today’s society<br />

is enormous and these products make an invaluable<br />

contribution to the quality of our lives. In medicine,<br />

the design and manufacture of pharmaceutical<br />

products has enabled us to cure diseases that<br />

have ravaged humankind throughout history. Crop<br />

protection and growth enhancement chemicals have<br />

enabled us to increase our food yields dramatically.<br />

It is particularly revealing to note that, although the<br />

twentieth century saw an increase in world population<br />

from 1.6 to 6 billion, it also saw an increase in<br />

life expectancy of almost 60% [1]!<br />

Chemistry has played, and continues to play, a<br />

fundamental role in almost every aspect of modern<br />

society, and, as the enormous populations in China,<br />

India and the emerging nations demand western<br />

levels of healthcare, food, shelter, transport and consumer<br />

goods, so the demands on the chemicals<br />

industries will grow.<br />

The successful development of the chemicals<br />

industries has almost had an inverse relationship<br />

with public perception. Since writing, over five years<br />

ago, in the introduction to The Chemistry of Waste<br />

Minimisation, that ‘The public image of the chemical<br />

industry has badly deteriorated in the last ten<br />

years . . .’ [2], the situation has worsened. Major surveys<br />

of public opinion throughout Europe in 2000<br />

revealed that in no country was the majority of<br />

people favourably disposed towards the chemical<br />

industry [3,4]. The most favourable interpretation of<br />

the data is that in some of the major centres of chemicals<br />

manufacturing (e.g. Germany) more people<br />

gave positive than negative views on chemicals<br />

JAMES H. CLARK<br />

1<br />

53<br />

Handbook of Green Chemistry and Technology<br />

Edited by James Clark, Duncan Macquarrie<br />

Copyright © 2002 by Blackwell Science Ltd<br />

manufacturing, but for many European countries<br />

the ratio of unfavourable to favourable views was<br />

alarmingly high (e.g. Sweden, 2.8; France, 2.2;<br />

Spain, 1.5; Belgium, 1.3).<br />

In the UK, a steady decline in public perception<br />

over many years is clearly evident (Fig. 1.1). It is<br />

especially disturbing to analyse the survey data more<br />

closely and to note, for example, that the 16–24-year<br />

age group has the lowest opinion of the chemicals<br />

industries. This is the most critical group for chemistry.<br />

We need to maintain a high level of interest<br />

and enthusiasm for chemistry at secondary and tertiary<br />

education levels so that we can maintain the<br />

supply of a large number of highly intelligent,<br />

motivated and qualified young people for our industries,<br />

universities, schools and other walks of life. At<br />

present, however, the poor image of chemistry is<br />

adversely affecting demand. In the UK, for example,<br />

the number of applicants to read chemistry at university<br />

has been falling steadily for several years<br />

(Fig. 1.2).<br />

The number of applicants to read chemical engineering<br />

is even more alarming (


2 Chapter 1<br />

60<br />

50favourable<br />

40<br />

30<br />

20<br />

10<br />

unfavourable<br />

1980 1990 2000<br />

Fig. 1.1 Trends in the favourability to the chemical industry of<br />

the general public (smoothed plots) (based on MORI Opinion<br />

Poll figures in the period 1980–2000).<br />

4000<br />

3500<br />

3000<br />

1996<br />

Fig. 1.2 Trend in the number of applications to study<br />

chemistry in UK universities (source: UCAS).<br />

2000<br />

foods, drinks or consumer products. It is revealing to<br />

note the recent change in name of the leading trade<br />

association for the chemicals industry in the USA<br />

from The American Chemical Manufacturers Association<br />

to The American Chemistry Council. Indeed, a<br />

cynical view might be that we can solve our image<br />

problems overnight by reinventing ourselves as<br />

‘molecular engineers’!<br />

In 1995 I wrote that chemistry’s bad image was<br />

‘. . . largely due to concerns over adverse environmental<br />

impact’ [2]. The growth in the chemicals<br />

54<br />

industries in the twentieth century was at the cost<br />

of producing millions of tonnes of waste, and if we<br />

extend the discussion to include health and safety<br />

issues then we must add the chemical disasters that<br />

have led to much unfavourable publicity and have<br />

hardened the views of many critics. The increasing<br />

levels of environmental awareness among the<br />

general public make it even more important that<br />

the chemicals industries ‘clean up their act’. Public<br />

acceptability of environmental pressure groups adds<br />

to their influence and together they effectively force<br />

governments to use legislation to force industry into<br />

making improvements.<br />

How much do we need to change? Although early<br />

work to ‘green’ the manufacture of chemicals was<br />

focused largely on reducing the environmental<br />

impact of chemical processes, a much wider view<br />

will be necessary in the new century. An exaggerated<br />

but illustrative view of twentieth century chemical<br />

manufacturing can be written as a recipe [5]:<br />

(1) Start with a petroleum-based feedstock.<br />

(2) Dissolve it in a solvent.<br />

(3) Add a reagent.<br />

(4) React to form an intermediate chemical.<br />

(5) Repeat (2)–(4) several times until the final<br />

product is obtained; discard all waste and spent<br />

reagent; recycle solvent where economically<br />

viable.<br />

(6) Transport the product worldwide, often for longterm<br />

storage.<br />

(7) Release the product into the ecosystem without<br />

proper evaluation of its long-term effects.<br />

The recipe for the twenty-first century will be very<br />

different:<br />

(1) Design the molecule to have minimal impact<br />

on the environment (short residence time,<br />

biodegradable).<br />

(2) Manufacture from a renewable feedstock (e.g.<br />

carbohydrate).<br />

(3) Use a long-life catalyst.<br />

(4) Use no solvent or a totally recyclable benign<br />

solvent.<br />

(5) Use the smallest possible number of steps in the<br />

synthesis.<br />

(6) Manufacture the product as required and as<br />

close as possible to where it is required.<br />

The broader picture will apply not only to chemical<br />

manufacturing but also to transportation, legislation


and, most critically, education. We must train the<br />

new generation of chemists to think of the environmental,<br />

social and economic factors in chemicals<br />

manufacturing.<br />

1.2 The costs of waste<br />

In the time taken to read one page of this book,<br />

several tonnes of hazardous waste will have been<br />

released to the air, water and land by industry, and<br />

the chemicals industry is by far the biggest source of<br />

such waste. This is only a fraction of the true scale<br />

of the problem. Substances classified as ‘hazardous’<br />

only represent a very small number of the total<br />

number of substances in commercial use. In the mid-<br />

1990s in the USA, for example, only about 300 or<br />

so of the 75000 commercial substances in use were<br />

classified as hazardous. Clearly a much higher proportion<br />

of commercial chemicals presents a threat to<br />

humans and to the environment, and as mounting<br />

pressure will lead to an ever-increasing number<br />

of chemicals being tested then the scale of the<br />

‘hazardous waste’ problem will take on ever more<br />

frightening proportions. Yet this only represents one<br />

‘cost’ of waste and the cost of waste can be truly<br />

enormous.<br />

Compliance with existing environmental laws will<br />

cost new EU member states well over E10 billion; a<br />

similar amount is spent each year in the USA to treat<br />

and dispose of waste. Governments across the globe<br />

are increasing the relative costs of waste disposal to<br />

discourage the production of waste and to encourage<br />

recycling and longer product lifetimes.<br />

Although, in general terms, company accounting<br />

practices are highly developed, when it comes to industrial<br />

chemical processes, particularly for smaller<br />

companies working with multi-purpose plants in the<br />

speciality chemicals area, the true breakdown of<br />

manufacturing costs is often unknown. Sophisticated<br />

process monitoring and information technology<br />

developments are beginning to allow the true<br />

production costs to become evident. What this shows<br />

is that the cost of waste can easily amount to 40%<br />

of the overall production costs for a typical speciality<br />

chemical product (Fig. 1.3).<br />

However, the costs of effluent treatment and waste<br />

disposal actually tell only part of the story. There are<br />

other direct costs to production resulting from inefficient<br />

manufacturing, by-product generation and<br />

raw material and energy inefficiencies. Industry also<br />

55<br />

Labour<br />

Capital<br />

depreciation<br />

Energy<br />

utilities<br />

Waste<br />

is becoming increasingly aware of the indirect costs<br />

of waste on deteriorating public relations (as<br />

described in Section 1.1). These affect the attitudes<br />

of the workforce and hence their morale and performance,<br />

and also that of their neighbours who can<br />

lobby local authorities to impose tighter standards<br />

and legislation. As a society, we can add the largely<br />

unknown but certain to be substantial (if not catastrophic)<br />

costs to the environment (including human<br />

health). All of these costs will grow into the future<br />

through tougher legislation, greater fines, increased<br />

waste disposal costs, greater public awareness and<br />

diminishing raw materials, forcing the adoption of<br />

more efficient manufacturing (Fig. 1.4).<br />

1.3 The greening of chemistry<br />

Introduction 3<br />

Materials<br />

Fig. 1.3 Production costs for speciality chemicals.<br />

Sustainable development is now accepted by governments,<br />

industry and the public as a necessary goal<br />

for achieving the desired combination of environmental,<br />

economic and societal objectives. The challenge<br />

for chemists and others is to develop new<br />

products, processes and services that achieve all the<br />

benefits of sustainable development. This requires a<br />

new approach whereby the materials and energy<br />

input to a process are minimised and thus utilised<br />

at maximum efficiency. The dispersion of harmful<br />

chemicals in the environment must be minimised or,<br />

preferably, completely eliminated. We must maxi-


4 Chapter 1<br />

Workforce Process waste<br />

Public<br />

relations<br />

Neighbours<br />

Raw<br />

material<br />

inefficiencies<br />

Effluent<br />

treatment<br />

COSTS OF WASTE<br />

Production losses<br />

Energy<br />

inefficiencies<br />

Waste<br />

disposal<br />

mise the use of renewable resources and extend the<br />

durability and recyclability of products, and all of this<br />

must be achieved in a way that provides economic<br />

benefit to the producer (to make the greener product<br />

and process economically attractive) and enables<br />

industry to meet the needs of society.<br />

We can start by considering the options for waste<br />

management within a chemical process (Fig. 1.5).<br />

The hierarchy of waste management techniques<br />

now has prevention, through the use of cleaner<br />

processes, as by far the most desirable option. Recycling<br />

is considered to be the next most favourable<br />

option and, from an environmental standpoint,<br />

is particularly important for products that do not dissipate<br />

rapidly and safely into the environment.<br />

Disposal is certainly the least desirable option. The<br />

term ‘cleaner production’ encompasses goals and<br />

principles that fall nicely within the remit of waste<br />

minimisation. The United Nations Environmental<br />

Programme describes cleaner production as:<br />

‘The continuous application of an integrated preventative<br />

environmental strategy to processes and<br />

products to reduce risks to humans and the environment.<br />

For production processes, cleaner production<br />

includes conserving raw materials, and<br />

Byproduct<br />

formation<br />

56<br />

Damage to<br />

environment<br />

Fig. 1.4 The costs of waste.<br />

reducing the quality and toxicity of all emissions<br />

and wastes before they leave a process.’<br />

Cleaner production and clean synthesis fall under<br />

the heading of waste reduction at source and, along<br />

with retrofitting, can be considered as the two<br />

principal technological changes. Waste reduction<br />

at source also covers good housekeeping, input<br />

material changes and product changes.<br />

There are many ways to define the efficiency of a<br />

chemical reaction. Yield and selectivity traditionally<br />

have been employed, although these do not necessarily<br />

give much information about the waste produced<br />

in a process. From an environmental (and<br />

increasingly economic) point of view, it is more<br />

important to know how many atoms of the starting<br />

material are converted to useful products and how<br />

many to waste. Atom economy is a quantitative<br />

measure of this by, for example, calculating the percentage<br />

of oxygen atoms that end up in the desired<br />

product [6]. We can illustrate this by considering<br />

a typical oxidation reaction whereby an alcohol,<br />

for example, is converted to a carboxylic acid using<br />

chromium(VI) as the stoichiometric oxidant. The<br />

material inputs for this reaction are the organic substrate,<br />

a source of chromium(VI), acid (normally sul-


Fig. 1.5 Options for waste<br />

management within a chemical<br />

manufacturing process.<br />

Good<br />

housekeeping<br />

Table 1.1 ‘Atom accounts’ for a typical partial oxidation<br />

reaction using chromate<br />

Retro-fitting<br />

Element Fate Atom utilisation<br />

C Product(s) Up to 100%<br />

H Product(s) + waste acid


6 Chapter 1<br />

Table 1.3 Global ‘lost work’ in major chemical processes<br />

Process Raw materials Final product a<br />

The term ‘green chemistry’ is becoming the worldwide<br />

term used to describe the development of<br />

more eco-friendly, sustainable chemical products<br />

and processes. The term was coined almost ten years<br />

ago by the US Environmental Protection Agency and<br />

has been defined as:<br />

‘The utilisation of a set of principles that reduces<br />

or eliminates the use or generation of hazardous<br />

substances in the design, manufacture and application<br />

of chemical products’ (Paul Anastas)<br />

This is elaborated further in the form of the so-called<br />

Principles of Green Chemistry:<br />

• Waste prevention is better than treatment or<br />

clean-up<br />

• Chemical synthesis should maximise the incorporation<br />

of all starting materials<br />

• Chemical synthesis ideally should use and generate<br />

non-hazardous substances<br />

• Chemical products should be designed to be nontoxic<br />

• Catalysts are superior to reagents<br />

• The use of auxiliaries should be minimised<br />

• Energy demands in chemical syntheses should be<br />

minimised<br />

• Raw materials increasingly should be renewable<br />

• Derivations should be minimised<br />

• Chemical products should break down into innocuous<br />

products<br />

• Chemical processes require better control<br />

• Substances should have minimum potential for<br />

accidents<br />

The chemical technologies, both new and established,<br />

that are described in this book address these<br />

principles by considering atom efficiency, alternative<br />

energy sources, the use of alternative feedstocks, in-<br />

Theoretical work potential (kJmol -1 final product)<br />

Natural gas + air Æ methanol 1136 717 63<br />

Natural gas + air Æ hydrogen 409 236 58<br />

Ammonia (from natural gas + air) Æ nitric acid 995 43 4<br />

Copper ore Æ copper 1537 130 9<br />

Bauxite Æ aluminium 4703 888 19<br />

a Excludes any ‘steam credit’.<br />

58<br />

2.0 x 10 9 ha<br />

for food<br />

production<br />

for 10 bn people<br />

2.8 x 10 9 ha available<br />

1 x 10 9 ta -1<br />

for organics<br />

Fig. 1.6 Biomass utilisation in 2040.<br />

Thermodynamic efficiency (%)<br />

0.8 x 10 9 ha<br />

available for<br />

non-foods<br />

@ 40 t ha -1 a -1<br />

32 x 10 9 ta -1<br />

+ forests<br />

+ waste streams<br />

50 x 10 9 ta -1<br />

of biomass<br />

>40 x 10 9 ta -1<br />

for energy<br />

= 2 x 10 20 Ja -1<br />

novative engineering, clean synthesis and process<br />

improvements.<br />

It is, perhaps, worth focusing briefly on one of<br />

these principles as we enter the century where oil<br />

reserves will be seriously diminished: ‘Raw materials<br />

increasingly should be renewable’. Can we base<br />

the future chemical industry on biomass? Remarkably,<br />

at least some of the better calculations show<br />

that this is a very likely scenario [10]. With a modest<br />

increase in farming efficiency to improve crop yield<br />

to about 40tha -1 year -1 we will need only less than<br />

1% of the biomass available globally to provide all<br />

the raw material necessary to feed the entire organic<br />

chemicals industry by 2040 (Fig. 1.6).


Assuming a global population of 10 billion by<br />

that year, we can still reasonably expect to produce<br />

enough ‘spare’ biomass to supply some 19% of the<br />

energy requirements of that future society (Table<br />

1.4).<br />

This would still make us very reliant on fossil fuels<br />

but, significantly, much less dependent on oils, the<br />

most vulnerable of the major energy sources based<br />

on the current rate of utilisation. Feeding, maintaining<br />

and providing material comforts for all is indeed<br />

within our grasp if, to paraphrase Mahatma Gandhi,<br />

we seek to satisfy our need and not our greed.<br />

By incorporating raw materials considerations into<br />

the ‘big picture’ we can move towards the ultimately<br />

Table 1.4 From fossil to green<br />

Energy source 1990 a<br />

Percentage of<br />

energy sources<br />

2040 b<br />

Oil 38 17<br />

Coal 20 18<br />

Gas 16 14<br />

Biomass 16 19<br />

Hydro 5 5<br />

Nuclear 5 6<br />

Solar — 14<br />

Wind — 7<br />

a Based on an energy consumption of 3.5 ¥ 10 20 J.<br />

b Based on an energy consumption of 1 ¥ 10 21 J.<br />

Fig. 1.7 Life-cycle assessment for<br />

chemical products (E = energy input;<br />

C = consumables input; W = waste).<br />

E<br />

Premanufacturing<br />

essential concept of life-cycle assessment. The lifecycle<br />

of a product can be considered as [11]:<br />

Pre-manufacturing (materials acquistition)<br />

Manufacturing (processing and formulation)<br />

Product delivery (packaging and distribution)<br />

Product use<br />

End of (first) life<br />

Introduction 7<br />

This model can be elaborated for a life-cycle assessment<br />

for chemical products (Fig. 1.7).<br />

This quickly allows us to recognise the vital importance<br />

of the other end of the product cycle: end of<br />

life. The recycling of waste is not embraced strictly<br />

by the principles of green chemistry, because they<br />

are focused on avoiding waste at source, but its<br />

importance cannot be ignored. If a product cannot<br />

be dissipated quickly and safely into the environment,<br />

then it is essential that it or its component<br />

parts are efficiently recycled. We can no longer afford<br />

single-use products.<br />

Pollution prevention options can be considered at<br />

every stage in the life-cycle of a chemical product<br />

(Fig. 1.8) [11]. In general, we should be looking<br />

C E C E E C<br />

E<br />

Manufacturing<br />

W W W W W<br />

59<br />

remanufacture<br />

recycle<br />

Product<br />

delivery<br />

Product<br />

use<br />

refurbish<br />

End of Life


8 Chapter 1<br />

Premanufacturing<br />

Alternative<br />

feedstocks/<br />

renewable<br />

resources<br />

Avoid<br />

auxiliaries<br />

Minimise<br />

transportation<br />

Utilise waste<br />

Minimise<br />

solvents<br />

Manufacturing<br />

Atom efficient<br />

Catalysts<br />

rather than<br />

reagents<br />

Solvent<br />

substitution<br />

Safer<br />

chemistry<br />

Simpler<br />

chemistry<br />

Minimum<br />

energy<br />

Avoid<br />

additives<br />

Fig. 1.8 Pollution prevention options in the life-cycle of a<br />

chemical product.<br />

increasingly at the industrial ecology goals for green<br />

chemistry [11]:<br />

• Adopt a life-cycle perspective regarding chemical<br />

products and processes<br />

• Realise that the activities of your suppliers and<br />

customers determine, in part, the greenness of<br />

your product<br />

• For non-dissipative products, consider recyclability<br />

• For dissipative products (e.g. pharmaceuticals,<br />

crop-protection chemicals), consider the environmental<br />

impact of product delivery<br />

• Perform green process design as well as green<br />

product design<br />

Life-cycle assessment is given better and more<br />

detailed consideration in later chapters in this book.<br />

A number of green chemistry and sustainable<br />

chemistry initiatives now are in place or becoming<br />

Product<br />

delivery<br />

Minimal<br />

transportation<br />

Minimal<br />

packaging/<br />

eco-friendly<br />

packaging<br />

60<br />

Product<br />

use<br />

Minimal<br />

consumption/<br />

maximum<br />

efficiency<br />

Minimal<br />

auxiliary<br />

needs<br />

Minimal<br />

energy usage<br />

End of life<br />

Biodegradable<br />

Recyclable<br />

Environmentally<br />

compatible<br />

established in various corners of the globe, including<br />

the USA, UK, Australia and Japan. In the UK, the<br />

Green Chemistry Network (GCN) [12] was established<br />

in 1998 with funding from the Royal Society<br />

of Chemistry. The GCN has its hub based at the University<br />

of York in England and benefits from close<br />

collaboration with the world-famous Science Education<br />

Group and the Chemical Industries Education<br />

Centre, as well as the staff of one of the UK’s most<br />

successful Chemistry Departments. The GCN promotes<br />

green chemistry through increased awareness,<br />

education and training and facilitates the sharing of<br />

good practice in green chemistry through conferences,<br />

technology transfer activities and by acting as<br />

a focal point for relevant information. It is doing this<br />

by providing educational material for all levels, training<br />

courses for industrialists and teachers, conferences<br />

and seminars on green chemistry, technology<br />

transfer brokerage, databases of green chemistry<br />

articles and links to other relevant activities, notably<br />

through a dynamic website.<br />

The GCN works alongside the Royal Society of


Chemistry journal, Green Chemistry, which provides<br />

news on grants, initiatives, educational and industrial<br />

development and conferences, as well as a<br />

who’s-who on green chemistry research and regular<br />

peer-reviewed articles from chemistry and chemical<br />

engineering university departments and chemical<br />

and pharmaceutical companies across the world.<br />

The US Green Chemistry Institute (GCI) has<br />

been promoting the principles and practice of green<br />

chemistry for several years. The GCI, which now has<br />

core funding from, and links with, the American<br />

Chemical Society (ACS), is dedicated to encouraging<br />

environmentally benign chemical synthesis and promoting<br />

research and education. The US Presidential<br />

Green Chemistry Awards Programme recognises and<br />

publicises achievements by industry and academe,<br />

encouraging industry to talk openly about its innovative<br />

clean chemistry and providing scientists and<br />

education with some excellent case studies. There<br />

are now green-chemistry-related award schemes<br />

in several other countries, including the UK, Italy,<br />

Australia and Japan. Major green chemistry or sustainable<br />

chemistry networks and related initiatives<br />

have been set up across the globe with significant<br />

developments, including a series of Gordon Green<br />

Chemistry conferences (in the USA and UK) and the<br />

61<br />

first International Union of Pure and Applied Chemistry<br />

(IU<strong>PAC</strong>) International Symposium of Green<br />

Chemistry (in India). The greening of chemistry is<br />

truly underway!<br />

References<br />

Introduction 9<br />

1. Breslau, R. Chemistry Today and Tomorrow. Awareness<br />

Chemical Society, Washington, 1997.<br />

2. Clark, J. H. The Chemistry of Waste Minimisation. Blackie<br />

Academic, London, 1995.<br />

3. MORI. The Public Image of the Chemical Industry. Research<br />

study conducted for the Chemical Industries<br />

Association. MORI, London, 1999.<br />

4. CEFIC. CEFIC Pan European Survey 2000. Image of the<br />

Chemical Industry Summary. CEFIC, Brussels, 2000.<br />

5. Based on: Woodhouse, E. J. Social Reconstruction of a<br />

Technoscience?: The Greening of Chemistry.<br />

http://www.rpi.edu/~woodhe/docs/green.html<br />

6. Trost, B. M. Angew Chem. Int. Food Engl., 1995, 34, 259.<br />

7. Clark, J. H. Green Chem., 1999, 1, 1.<br />

8. Sheldon, R. A. Chemtech, 1994, March, 38.<br />

9. Hinderink, A. P., van der Kooi, H. J., & de Swaan<br />

Arons, J. Green Chem., 1999, 1, G176.<br />

10. Okkerse, C., & van Bekkum, H. Green Chem., 1999, 1,<br />

107.<br />

11. Anastas, P. T., & Lankey, R. L. Green Chem., 2000, 2,<br />

289.<br />

12. See http://www.rsc.org/greenchem/


1 Introduction<br />

Chapter 2: Principles of Sustainable<br />

and Green Chemistry<br />

In the modern context, the terms ‘sustainable development’<br />

and ‘green chemistry’ have been around for<br />

less than 15 years. Discussion of sustainability began,<br />

essentially, when the 1987 UN Commission on Environment<br />

and Development (usually referred to as<br />

the Bruntland Commission) noted that economic<br />

development might lead to a deterioration, not an<br />

improvement, in the quality of people’s lives [1].<br />

This led to the now commonly accepted definition of<br />

‘sustainable development’ as being:<br />

‘Development which meets the needs of the<br />

present without compromising the ability of future<br />

generations to meet their own needs.’<br />

This definition is intentionally broad, covering all<br />

aspects of society. The debate on what it actually<br />

means, in practical terms, for different disciplines<br />

and sectors of society continues, and indeed there<br />

are those who argue that it is a contradiction in<br />

terms. However, working interpretations of the definition<br />

are becoming established [2]. For example, in<br />

planning it is the process of urban revitalisation that<br />

seeks to integrate urbanisation with nature preservation;<br />

in biology it is associated with the protection<br />

of biodiversity; in economics it is the accounting for<br />

‘natural capital’.<br />

Sustainable development has particular relevance<br />

for chemistry-based industries because it is concerned<br />

with avoidance of pollution and the reckless<br />

use of natural resources. In essence it is being recognised<br />

increasingly as the pursuit of the principles and<br />

goals of green chemistry.<br />

The Green Chemistry Movement was started in<br />

the early 1990s by the US Environmental Protection<br />

Agency (EPA) as a means of encouraging industry<br />

and academia to use chemistry for pollution prevention.<br />

More specifically, the green chemistry mission<br />

was:<br />

‘To promote innovative chemical technologies that<br />

reduce or eliminate the use or generation of haz-<br />

MIKE LANCASTER<br />

10<br />

62<br />

Handbook of Green Chemistry and Technology<br />

Edited by James Clark, Duncan Macquarrie<br />

Copyright © 2002 by Blackwell Science Ltd<br />

ardous substances in the design, manufacture and<br />

use of chemical products.’<br />

In conjunction with the American Chemical Society,<br />

the EPA developed green chemistry into a set of<br />

twelve guiding principles [3]. These principles can be<br />

summarised as being concerned with ensuring that:<br />

• The maximum amounts of reagents are converted<br />

to useful product (atom economy)<br />

• Production of waste is minimised through reaction<br />

design<br />

• Non-hazardous raw materials and products are<br />

used and produced wherever possible<br />

• Processes are designed to be inherently safe<br />

• Greater consideration is given to using renewable<br />

feedstocks<br />

• Processes are designed to be energy efficient<br />

These principles and associated terminology are<br />

becoming widely accepted as a universal code of<br />

practice as the Green Chemistry Movement spreads<br />

out of the USA into Europe, Australia and Asia. It is<br />

evident from these principles that green chemistry<br />

encompasses much more of the concepts of sustainability<br />

than simply preventing pollution; two important<br />

aspects are the design for energy efficiency and<br />

the use of renewable feedstocks.<br />

This chapter will explore some of the key features<br />

of these principles, many of which will be dealt with<br />

in greater depth in other parts of the book, and assess<br />

the relevance and opportunities for the chemical<br />

industry.<br />

2 Green Chemistry and Industry<br />

Chemical companies worldwide now are taking the<br />

issue of sustainable and green chemistry seriously.<br />

A combination of increasing amounts of legislation,<br />

increased public awareness and concern and the<br />

realisation that eco-efficiency is good for business are<br />

rapidly increasing the rate of change. The first real<br />

proof that the chemical industry was serious about<br />

environmental concerns came with the ‘responsible


Table 2.1 Key principles and indicators<br />

of the Responsible Care Programme<br />

care’ concept, which was developed by the Canadian<br />

Chemical Producers Association in 1989 and has<br />

been adopted since by many industry association<br />

members throughout the world. The key behind<br />

responsible care is the continuous delivery of health,<br />

safety and environmental improvements related to<br />

both products and processes. Some of the guiding<br />

principles and performance indicators of the Responsible<br />

Care Programme are shown in Table 2.1; the<br />

similarity of many of these to the principles of green<br />

chemistry is self-evident.<br />

Despite this overall commitment, environmental<br />

protection also is often seen by industry as a necessary<br />

cost to comply with increasingly stringent legislation.<br />

There is a great deal of justification in this<br />

view; a recent survey, commissioned by the UK<br />

Department of Environment Transport and Regions<br />

[4], showed that expenditure on environmental<br />

protection, by the UK industry, had risen from £2482<br />

million in 1994 to £4274 million in 1997. The chemical<br />

industry bore the brunt of this expenditure,<br />

which was some 24% of the total spend. Looking<br />

at the capital expenditure element of these figures<br />

(Fig. 2.1), it is evident that the chemical industry<br />

is heavily focused on end-of-pipe solutions rather<br />

than on the integrated process approach, which<br />

would prevent many of the environmental issues<br />

arising.<br />

Clearly there is significant scope for wider adoption<br />

and investment in cleaner and greener<br />

processes, thus avoiding the need for much of the<br />

end-of-pipe expenditure. One of the major goals of<br />

green chemistry is to demonstrate that adoption of<br />

the principles, by industry, can create a competitive<br />

advantage [5]. In this context it is helpful to look at<br />

green chemistry as a reduction process.<br />

Principles of Sustainable and Green Chemistry 11<br />

Principles Indicators<br />

Resource Conservation—waste reduction Safety—lost time accidents<br />

Experience Learning—sharing best practice Waste Emissions—continuous<br />

reduction<br />

Process Safety—risk management Energy Consumption—targeted<br />

improvements<br />

Product Stewardship—risk assessment<br />

Policy—HSE policy to reflect commitment<br />

Management Systems—address impact of activities<br />

HSE, Health & Safety Executive.<br />

63<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

End of Pipe<br />

capex<br />

Integrated<br />

Process capex<br />

Fig. 2.1 The 1997 capital expenditure by the UK chemical<br />

industry (£million) on environmental protection.<br />

Risk and<br />

hazard<br />

Cost Waste<br />

REDUCING<br />

Energy<br />

Fig. 2.2 Green chemistry as a reduction process.<br />

Materials<br />

The simple model of Fig. 2.2 incorporates the key<br />

elements of green chemistry in a way that finance<br />

directors, environmentalists, production managers,<br />

R&D technologists and chief executives can all<br />

understand and, hopefully, buy into.<br />

By looking at the principles of green chemistry as<br />

a tool-kit for achieving this reduction process, it


12 Chapter 2<br />

becomes more evident that as waste, energy, etc. are<br />

reduced the cost of the process also normally will be<br />

reduced. This economic advantage undoubtedly will<br />

be the biggest driver for change. There will, however,<br />

be other advantages for industry, not least of which<br />

will be an improvement in the public image, which<br />

is at an all time low in many countries [6] mainly<br />

due to the perception that the industry is environmentally<br />

unfriendly. We can see now how green<br />

chemistry becomes connected to the increasingly<br />

important business concept of the ‘triple bottom line’<br />

in which business performance is measured not only<br />

in terms of profitability but also in terms of the environmental<br />

and social performance of the company.<br />

Although it is easy to buy into the concepts of sustainable<br />

development, it is often more difficult to<br />

achieve the objectives in practice. Many of these difficulties<br />

are to do with culture and the way chemistry<br />

and related disciplines are taught and practised.<br />

What is really required is a culture change both in<br />

education and industry. In education the principles<br />

of green chemistry need to be the underlying theme,<br />

not taught in isolation. In industry the principles<br />

of sustainability should form part of the company<br />

ethos and be reflected in management systems and<br />

procedures.<br />

3 Waste Minimisation and Atom Economy<br />

3.1 Atom economy<br />

Generations of chemists, especially organic chemists,<br />

have been educated to devise synthetic reactions<br />

to maximise yield and purity. Although these are<br />

worthy goals, reactions may proceed in 100% yield<br />

to give a product of 100% purity and still produce<br />

more waste than product. In simplistic terms, Equation<br />

2.1:<br />

SO 3 Na<br />

+ 2NaOH<br />

ONa<br />

A + B Æ C + D + E (2.1)<br />

in which A and B react to give product C in high<br />

yield and high purity, also leads to the formation of<br />

by-products (or waste) D and E in stoichiometric<br />

quantities. For many years phenol was manufactured<br />

via the reaction of sodium benzene sulfonate<br />

(from benzene sulfonation) with sodium hydroxide;<br />

the products of this reaction are sodium phenolate<br />

(which is hydrolysed subsequently to phenol),<br />

sodium sulfite and water. Even if the reaction proceeds<br />

in quantitative yield, it is evident from looking<br />

at the molecular weights of the product (sodium<br />

phenolate) and unwanted by-products (sodium<br />

sulfite and water) that, in terms of weight, the reaction<br />

produces more waste than product. Historically,<br />

however, the chemist would not consider the production<br />

of this aqueous salt waste to be of any<br />

importance when designing the process.<br />

The atom economy concept proposed by Trost [7]<br />

is one of the most useful tools available for design of<br />

reactions with minimum waste. The concept is that<br />

for economic and environmental reasons reactions<br />

should be designed to be atom efficient, i.e. as many<br />

of the reacting atoms as possible should end up in<br />

useful products. In the example shown in Fig. 2.3 all<br />

the carbon atoms present in the starting material are<br />

incorporated into the product, giving a carbon atom<br />

efficiency of 100%, but none of the sulfur ends up<br />

as useful product and hence the atom efficiency for<br />

sulfur is 0%. Overall, the atom efficiency of the reaction<br />

is defined as the ratio of the molecular weights<br />

of desired product to the sum of the molecular<br />

weights of all materials produced in the process. In<br />

the above example the atom efficiency would be<br />

116/260 or 44.6%.<br />

The concept of atom economy has been expanded<br />

usefully by Sheldon [8,9], by the introduction of the<br />

term ‘E factor’, which is the ratio of the kilograms of<br />

+ Na 2 SO 3 + H 2 O<br />

MWts 180 2 ¥ 40 116 126 18 Fig. 2.3 Benzene sulfonate route to<br />

phenol.<br />

64


y-product per kilogram of product. In this context,<br />

by-product is taken to mean everything other than<br />

useful product, including any solvent consumed.<br />

This concept is particularly useful for comparing<br />

industrial processes, where the yield also can be<br />

taken into account. Assuming a 100% yield for the<br />

example in Fig. 2.3, we would have an E factor of<br />

144/116 or 1.24 (the water produced could be justifiably<br />

ignored, improving the E factor to 1.08).<br />

According to Sheldon, this is typical of a bulk chemicals<br />

production process, however in the fine chemicals<br />

industry the E factor can be as high as 50<br />

whereas in the pharmaceuticals sector it may be<br />

even higher, which is striking evidence of the waste<br />

problem faced by chemistry-based industries.<br />

The cost and associated environmental problems<br />

of disposing the sodium sulfite produced in the<br />

phenol process contributed to its replacement, on<br />

economic grounds, by the cumene process. In this<br />

process the final step involves the acid-catalysed<br />

decomposition of cumene hydroperoxide to phenol<br />

and acetone (Fig. 2.4). In this case both the phenol<br />

and acetone are wanted products, hence apart from<br />

a very small amount of acid (used to aid decomposition<br />

of the hydroperoxide) this reaction has a<br />

100% atom efficiency and a zero E factor, indicating<br />

a completely waste-free process (assuming 100%<br />

yield).<br />

Of course the atom economy concept should not<br />

replace consideration of yield, ease of product isolation,<br />

purity requirements, etc. when devising a<br />

chemical synthesis but it should be thought of as an<br />

additional consideration. The economics of chemical<br />

production are changing, particularly in the fine,<br />

speciality and pharmaceuticals sectors, where waste<br />

generation and other environmental considerations<br />

are becoming an increasingly significant [10] proportion<br />

of the overall manufacturing cost.<br />

H3C CH3 OOH OH<br />

Fig. 2.4 Cumene route to phenol.<br />

+ CH 3COCH 3<br />

Principles of Sustainable and Green Chemistry 13<br />

65<br />

3.2 Some inherently atom economic reactions<br />

By their very nature some reaction types are likely<br />

to produce less waste than others by virtue of being<br />

inherently atom efficient. These reaction types are<br />

worth considering when devising a synthetic strategy.<br />

Obviously other factors also need to be taken<br />

into account in determining the most efficient, competitive<br />

and eco-friendly route. These factors<br />

include:<br />

• Cost and availability of raw materials<br />

• Toxicity/hazardous nature of raw materials<br />

• Yield<br />

• Ease of product isolation and purification<br />

• Solvent requirements<br />

• Energy requirements<br />

• Equipment requirements, cost and availability<br />

• Process times<br />

• Nature of waste materials<br />

Pericyclic [11] reactions occur via a concerted<br />

process through a cyclic transition state. These reactions<br />

are typified by the Diels–Alder reaction, sigmatropic<br />

rearrangements and cheletropic additions,<br />

amongst others, and have theoretical atom efficiencies<br />

of 100% and hence should be high on the<br />

list of considerations when designing synthetic<br />

pathways.<br />

Organic fungicides have played a vital role in<br />

ensuring the plentiful supply of food and, although<br />

many are not perfect from an environmental point<br />

of view, they are generally more eco-friendly and<br />

less toxic than the mercury-based fungicides that<br />

they replaced. One of the major classes of contact<br />

fungicides used today are the sulfenamides, typified<br />

by Captan and the related materials Folpet and<br />

Difoltan [12]. The starting reaction to all these materials<br />

is the Diels–Alder addition of maleic anhydride<br />

to butadiene (Fig. 2.5).<br />

Many Diels–Alder reactions are carried out in<br />

organic solvents, and non-recoverable lewis acids<br />

such as aluminium chloride frequently are used to<br />

extend the range and speed up the reactions. Both<br />

of these may detract from the ‘greenness’ but there<br />

are examples where these reactions occur rapidly<br />

without the use of catalysts or organic solvents.<br />

2,2,5-Trisubstituted tetrahydrofurans are a novel<br />

class of antifungal compounds; their synthesis<br />

involves the key Diels–Alder reaction shown in Fig.<br />

2.6. Saksena [13] found that the reaction proceeded


14 Chapter 2<br />

in virtually quantitative yield when water was used<br />

as solvent, whereas in organic solvents a highly complex<br />

mixture of products was obtained.<br />

Claisen rearrangements are another class of<br />

‘waste-free’ pericyclic reactions of significant importance.<br />

Claisen rearrangement of the propargyl ether,<br />

which proceeds in 85% yield (Fig. 2.7), is at the<br />

heart of a simple route to cordiachromene [14], a<br />

natural product that shows significant antibacterial<br />

and anti-inflammatory properties.<br />

Addition reactions, in which two molecules<br />

combine to form a single molecule of product, are<br />

another class of inherently waste-free reactions of<br />

66<br />

Fig. 2.5 Captan synthesis.<br />

Fig. 2.6 Water-based Diels–Alder<br />

reaction—intermediate to antifungal<br />

compounds.<br />

Fig. 2.7 Claisen rearrangement en<br />

route to cordiachromene.<br />

Fig. 2.8 A green Michael addition.<br />

widespread applicability. By their very nature, addition<br />

reactions involve an unsaturated bond and are<br />

typified by reactions such as electrophillic addition of<br />

halogens to alkenes and nucleophilic additions to<br />

carbonyls.<br />

Michael additions normally are carried out with<br />

base catalysts, however there have been several<br />

recent examples of very green Michael additions<br />

that occur in high yield in water. A striking example<br />

of this is the addition of acrolein to 2-methyl-1,<br />

3-cyclopentadione [15], which proceeds in quantitative<br />

yield in water at ambient temperature (Fig.<br />

2.8).


3.3 Some inherently atom<br />

uneconomic reactions<br />

Similarly there are reactions that will usually<br />

produce some waste material; these are typified by<br />

substitution and elimination reactions. These reactions<br />

should be viewed with caution when designing<br />

green syntheses and, if a viable alternative is not<br />

possible, attempts made either to recycle or to find a<br />

use for the eliminated or substituted product.<br />

The Wittig reaction is highly useful for forming<br />

carbon–carbon double bonds and is widely used<br />

industrially in the manufacture of vitamins and<br />

pharmaceuticals. Although normally proceeding in<br />

high yield under mild conditions, it is an inherently<br />

wasteful reaction producing a mole equivalent of<br />

phosphine oxide per mole of product (Fig. 2.9).<br />

The phosphine oxide normally is converted to<br />

calcium phosphate for disposal. It is this 0% phosphorus<br />

atom efficiency that makes the Wittig<br />

reaction expensive, as well as environmentally problematic,<br />

and limits its usefulness to the production of<br />

high-value-added products. The greenness of the<br />

reaction can be improved, however, by converting<br />

the oxide back to triphenyl phosphine [16,17]. This<br />

recycling process, developed by the multinational<br />

chemical company BASF, involves chlorination of<br />

the phosphine oxide with phosgene, reduction with<br />

aluminium powder and hydrolysis. Although not a<br />

particularly green process (because it involves the<br />

use of hazardous reagents and produces aluminium<br />

hydroxide waste), overall, comparing the whole<br />

processes including triphenyl phosphine manufacture<br />

(Equation 2.2), the BASF route is more environmentally<br />

benign and cost effective.<br />

PCl + 3CHCl + 6Na Æ P( C H ) + 6NaCl<br />

3 6 5 6<br />

(2.2)<br />

Specific, more environmentally benign alternatives<br />

to the Wittig reaction now are being sought. The key<br />

(intermediate 1) to the anti-HIV drug Efavirenza has<br />

been produced in a one-pot process (see Scheme 2.1)<br />

with an overall yield of 92% [18]. The process<br />

involves reaction of cyclopropylcarboxaldehyde (the<br />

Fig. 2.9 Typical Wittig reaction.<br />

5 3<br />

Principles of Sustainable and Green Chemistry 15<br />

67<br />

same starting material as used in the Wittig reaction)<br />

with trichloromethyl anion generated in situ, acetylation<br />

and removal of acetate and chloride groups.<br />

The process still produces significant amounts of<br />

waste but it is much more environmentally benign<br />

waste.<br />

4 Reduction of Materials Use<br />

Frequently, many chemical reactions involve the use<br />

of reagents such as protecting groups and so-called<br />

catalysts that do not end up in the useful product.<br />

Organic solvents, often thought to be essential but<br />

sometimes not actually required at all, fall into this<br />

category. Some of these materials end up as waste<br />

and some are recovered, but in all cases valuable<br />

resources and energy are consumed that do not form<br />

part of the required product.<br />

Materials and money often are wasted in the<br />

design of chemical reactors, and new thinking about<br />

plant and ancillary equipment design (the process<br />

intensification concept) is part of the chemical engineering<br />

solution to greener chemical processes.<br />

There is a significant amount of synergy between<br />

the chemistry and engineering approaches to materials<br />

reduction. Frequently, low reactor utilisation,<br />

because of large solvent volumes for example, may<br />

necessitate the building of additional plant. By using<br />

the concepts of green chemistry to integrate the<br />

Scheme 2.1


16 Chapter 2<br />

chemistry and plant design, significant material<br />

savings can be made.<br />

4.1 Catalytic solutions<br />

Organic chemists very rarely write balanced equations<br />

and this can hide a multitude of sins. Taking<br />

Friedel–Crafts reactions as an example, alkylation<br />

and acylation reactions often are referred to as being<br />

catalysed by lewis acids such as aluminium chloride;<br />

although this is partially true, it is frequently ignored<br />

that the acylation reaction requires more than stoichiometric<br />

amounts of AlCl3 (Fig. 2.10) [19].<br />

In the alkylation reaction AlCl3 is required only in<br />

small amounts, but in the acylation reaction it complexes<br />

with the ketone product and is taken out of<br />

the catalytic cycle. In both cases, reactions usually<br />

are quenched with water, leading to copious<br />

amounts of aluminous waste and releasing three<br />

equivalents of HCl. In the case of the acylation of<br />

1,3-dimethylbenzene and assuming a quantitative<br />

yield, more than 0.9kg of AlCl3 are wasted per kilogram<br />

of dimethyl acetophenone produced.<br />

A recent, but classic, example of overcoming the<br />

wasted materials issue in aromatic acylations is the<br />

Hoechst Celanese route to the analgesic ibuprofen<br />

[20]. The reaction involves acylation of isobutylbenzene<br />

with acetic anhydride, a process that had been<br />

carried out traditionally with AlCl3 in an organic<br />

solvent. The Hoechst Celanese process employs<br />

liquid HF as both a true catalyst and solvent, the HF<br />

68<br />

being, for all practical purposes, completely recycled.<br />

Although the purist could argue that this process<br />

may not be particularly ‘intrinsically safe’ due to the<br />

potential hazard associated with handling HF, it is<br />

considerably greener in the context of reduced materials<br />

consumption.<br />

Much academic and industrial research effort has<br />

gone into the greening of Friedel–Crafts processes,<br />

with the aim of developing benign, easily recyclable,<br />

inexpensive, active solid catalysts that are highly<br />

selective and avoiding wasted raw materials and byproducts.<br />

For the alkylation reaction zeolites generally<br />

have provided the commercial solution to this<br />

problem, with major areas of research work centred<br />

on the avoidance of olefin oligomerisation and the<br />

development of catalysts stable to the operating conditions.<br />

Returning to the production of phenol from<br />

cumene (see above), the first step involves the alkylation<br />

of benzene with propene, which was carried<br />

out originally with AlCl3; today, several commercial<br />

zeolite-based processes have been developed. The<br />

Mobil process, developed in 1993, employs a highsilica<br />

catalyst ZMS-5 that gives almost stoichiometric<br />

yields, whereas Dow Chemical have developed a<br />

process based on de-aluminated mordøenite [21].<br />

The development of solid acid catalysts to solve the<br />

many problems associated with the acylation reaction<br />

generally has proved more problematic, but for<br />

activated substrates such as aryl ethers Rhone-<br />

Poulenc have developed an H-beta-zeolite catalyst<br />

[22].<br />

Fig. 2.10 Typical Friedel–Crafts<br />

reactions.


Another very important industrial process that<br />

essentially gives a free ride to a reactant is that of<br />

aromatic nitration. Aromatic nitro compounds are<br />

used widely as intermediates for dyes, plastics and<br />

pharmaceuticals, and for monosubstituted aromatic<br />

substrates it is often the para-isomer that is the<br />

required product. Conventional nitration technology<br />

uses a mixture of concentrated nitric and sulfuric<br />

acids, the latter acid often being used in considerable<br />

molar excess. The sulfuric acid is present in order<br />

to generate nitronium ions, which are the active<br />

nitration species and, in principle, are still present<br />

unchanged in the product mix. In practice, the reaction<br />

mix usually is quenched with water, leading<br />

to copious amounts of acidic waste to be disposed<br />

of. Smith et al. [23] have developed a more selective<br />

para-alkylation procedure that does not involve the<br />

use of sulfuric acid. Para-selectivity is enhanced<br />

by the use of recoverable zeolites but more than<br />

equimolar amounts of acetic anhydride are required<br />

to generate the active nitrating species (CH3CO 2NO 2)<br />

and to mop up the water formed; material usage<br />

therefore is still high.<br />

True catalytic nitration technology has been<br />

developed using lanthanide(III) triflates [24]. Lanthanide(III)<br />

triflates are unusual in that they function<br />

as strong Lewis acids, are stable to water and<br />

hence are recoverable from aqueous solutions. Using<br />

ytterbium or scandium triflate at levels as low as 1<br />

mol.% and equimolar amounts of nitric acid, nitration<br />

of a range of aromatic compounds was achieved<br />

at around 90% conversion.<br />

Rearrangements inherently should be atom efficient<br />

processes but sometimes the ‘catalyst’ required<br />

to cause the rearrangement cannot be readily recovered<br />

and reused. This is the case with some<br />

production processes for ethylidene norbornene<br />

(ENB) from vinylidene norbornene (VNB) (Fig.<br />

2.11). The ENB is used as the ‘diene’ component in<br />

ethene–propene diene monomer (EPDM) rubbers<br />

and it is often manufactured by Diels–Alder reaction<br />

of cyclopentadiene with butediene, followed by<br />

rearrangement of the so-formed VNB using sodium/<br />

potassium amalgam in liquid ammonia. Although<br />

most of the liquid ammonia (which is also used as<br />

a solvent) is recovered, there is significant loss of<br />

metals. Sumitomo [25] have developed an alternative<br />

solid base catalyst (Na/NaOH on g-alumina) that<br />

avoids waste and improves the safety aspects of the<br />

process.<br />

Principles of Sustainable and Green Chemistry 17<br />

69<br />

Fig. 2.11 Rearrangement of vinylidene norbornene (VNB) to<br />

ethylidene norbornene (ENB).<br />

4.2 Question the need for protection<br />

Another major source of raw material wastage comes<br />

from the use of protecting groups, frequently used in<br />

the synthesis of pharmaceuticals; these are necessarily<br />

used in stoichiometric amounts. Not only are<br />

the raw materials wasted but their use frequently<br />

requires an additional two process steps, involving<br />

increased uses of solvents, lower yields, etc. Wherever<br />

possible, the use of ancillary reagents such<br />

as protecting groups should be avoided. An excellent<br />

example of process simplification in which a threestep<br />

route has been reduced to a single step by<br />

a biotransformation is the manufacture of 6aminopenicillanic<br />

acid, an antibiotic intermediate<br />

[26].<br />

The original process involved protection of the carboxylate<br />

group in penicillin G by silylation; this reaction<br />

also requires dimethyl aniline to remove the HCl<br />

produced during silylation (Fig. 2.12). In the biocatalytic<br />

process, genetically engineered and immobilised<br />

penicillin amidase is used to deacylate<br />

penicillin G directly.<br />

There are many additional green benefits to the<br />

biocatalytic process, including:<br />

• Avoidance of dichloromethane solvent—water is<br />

used in the biocatalytic process<br />

• Energy savings—reaction carried out at 30°C as<br />

against -50°C for the protection step<br />

• Fewer safety problems—PCl5 also was used in the<br />

non-biocatalytic process<br />

4.3 Reduction of non-renewable raw<br />

material use<br />

The debate on when the supply of crude oil and gas<br />

will run out will not be settled for some considerable<br />

time. There is, however a growing consensus of<br />

opinion that, at least as far as oil is concerned, if we


18 Chapter 2<br />

continue to use resources at the current rate we will<br />

face a significant shortage (combined with a very<br />

high price) some time in the second half of this<br />

century [27]. The use of non-renewable resources<br />

for chemicals manufacture must be put into perspective:<br />

approximately 90% of crude oil currently<br />

is used to provide energy via burning of oil, gasoline<br />

and diesel, with only 8% of crude being converted<br />

into chemicals. The two main arguments for reducing<br />

our dependency on fossils and increasing our use<br />

of renewable feedstocks are:<br />

(1) To conserve valuable supplies of fossil fuels<br />

for future generations (a core principle of<br />

sustainability).<br />

(2) To reduce global emissions of greenhouse gases,<br />

especially carbon dioxide (renewable resources<br />

being CO2-neutral overall).<br />

Reduction in the use of fossil fuels for chemicals<br />

manufacture will have some benefit on conserving<br />

resources and reducing CO2 emissions, but these will<br />

be small compared to what can be achieved by using<br />

renewable resources for energy production. Chemicals<br />

manufacture from renewable resources, therefore,<br />

ideally should provide additional benefits such<br />

as reduced hazard, more efficient process, reduced<br />

cost, reduced pollution, meeting market needs, etc.<br />

Additionally, it is important to look at the whole<br />

process, including growing, transport, etc., to ensure<br />

that the total energy consumed (or total CO2 emission)<br />

is lower when employing the renewable<br />

resource. Chemistry does have a vital role to play in<br />

reducing the requirement for fossil fuels, e.g. more<br />

efficient combustion processes, the development of<br />

energy-efficient solar and fuel cells and the production<br />

of biodiesel.<br />

70<br />

Fig. 2.12 Routes to 6-aminopenicillanic<br />

acid.<br />

Table 2.2 Some disadvantages of vegetable-oil-based diesel<br />

High viscosity<br />

Lower volatility<br />

Reactivity of unsaturated chains, leading to gum formation<br />

Increased coking<br />

There is nothing particularly new about using<br />

vegetable-based diesel oils [28] but during the last<br />

60 years or so the advent of relatively inexpensive<br />

and technically superior petroleum-based diesel has<br />

prevented their widespread use. In recent years there<br />

has been considerable renewed interest in the competitive<br />

production of biodiesel to overcome many<br />

of the environmental issues associated with the<br />

petroleum-based material [29]. Some of the disadvantages<br />

of vegetable-oil-based diesel are shown in<br />

Table 2.2.<br />

These disadvantages generally preclude the use of<br />

unmodified vegetable oils, although there are many<br />

examples of 20–50% blends with conventional diesel<br />

being used for prolonged periods [30]. Transesterification<br />

has been the major technique<br />

employed to overcome these technical problems<br />

(especially high viscosity), although at added cost.<br />

Typically the anhydrous oil (triglycerides) is heated<br />

with methanol and a basic catalyst to give a mixture<br />

of methyl esters and glycerol, which is recovered as<br />

a valuable co-product. Although sodium hydroxide<br />

and sodium methoxide are widely used as catalysts,<br />

a ‘green’ process involving a reusable immobilised<br />

lipase catalyst and supercritical carbon dioxide has<br />

been demonstrated [31].<br />

The main obstacle to widespread use of biodiesel<br />

is the cost, of which up to 75% can be the raw


Fig. 2.13 Polylactic acid synthesis.<br />

vegetable oil cost; this has focused attention onto the<br />

use of used cooking oil for example, but non-uniformity,<br />

availability and collection issues have prevented<br />

commercial use to date.<br />

Use of renewable resources for making polymers<br />

is an area receiving much attention due to the relative<br />

ease of making biodegradable plastics with<br />

useful chemical and physical properties. It is important<br />

to caution against the perception, however, that<br />

just because a plastic is made from a renewable<br />

resource it is automatically greener than one made<br />

from petroleum. Many petroleum-based polymers<br />

such as polyethylene [32] and polyisoprene are fairly<br />

readily biodegraded; it is the additives (antioxidants)<br />

specifically added to prevent degradation, thus<br />

ensuring a useful life, that are the causes of many<br />

of the environmental problems. As in the case of<br />

biodiesel, one of the main issues preventing growth<br />

of the ‘renewable polymers’ sector is cost. In many<br />

cases the cost is associated with the relatively small<br />

amount of the required chemical being present in<br />

the crop, entailing high extraction cost and the production<br />

of large quantities of waste. In these cases a<br />

holistic approach is required with, for example,<br />

waste biomass being used as a fuel.<br />

Recent advances in producing polylactic acid<br />

(PLA) from corn starch have lead to the building of<br />

the first large-scale commercial production unit by<br />

Cargill-Dow [33]. The commercial viability of the<br />

polymer relies on novel processing that can be used<br />

to manipulate the molecular weight, crystallinity and<br />

chain branching, enabling materials with a wide<br />

range of end uses and markets to be made. Potential<br />

applications for PLA include:<br />

• Packaging—PLA can have the processability of<br />

polystyrene and the strength properties of<br />

poly(ethylene terephthalate), with good resistance<br />

to fats and oils.<br />

• Textiles—PLA has good drape, wrinkle and UVlight-resistance<br />

properties.<br />

Principles of Sustainable and Green Chemistry 19<br />

71<br />

The process involves fermentation of unrefined dextrose,<br />

derived from corn, to give D- and L-lactic acids,<br />

which are converted to D-, L- and meso-lactides<br />

before polymerisation (Fig. 2.13). By controlling the<br />

D, L and meso ratio, together with molecular weight,<br />

polymer properties can be tailored to meet product<br />

specifications [34].<br />

Society in the not too distant future will need<br />

to find viable alternatives to the use of fossil fuels<br />

for energy and probably for the synthesis of many<br />

chemicals. If the solution is to grow our energy, as<br />

opposed to using solar cells for example, then we will<br />

need to face the issues concerned with land usage<br />

[35]. Although there is no real shortage of land on<br />

the planet, there are serious debates as to the viability<br />

of growing most of our energy needs. These<br />

debates centre on land quality, accessibility, nearness<br />

to population centres, etc.<br />

4.4 Process intensification<br />

When designing a chemical process the engineering<br />

aspects are as important as the chemistry and it is<br />

often the lack of interaction between chemists and<br />

engineers at an early enough stage that results in<br />

processes being developed that are not as green or<br />

efficient as they otherwise could be. In many ways<br />

process intensification can be regarded as the engineering<br />

solution to green chemistry problems;<br />

the concept originated in the 1970s as a means of<br />

making large reductions in the cost of processing<br />

systems [36]. Like many cost reduction concepts,<br />

process intensification is concerned mainly with<br />

reducing materials use and energy consumption by<br />

reducing plant footprint and increasing throughput.<br />

Some of the key aspects of process intensification are<br />

shown in Fig. 2.14 [37].<br />

A fuller account of process intensification is presented<br />

elsewhere in the book but in the context<br />

of materials reduction it is worth mentioning an


20 Chapter 2<br />

Reaction Non-reaction<br />

Spinning disc reactor<br />

Supercritical fluids<br />

Static mixer reactor<br />

Reactor<br />

Static mixing catalysts<br />

Operation<br />

Monolithic reactors<br />

Microreactors<br />

Heat exchange<br />

reactors<br />

Supersonic gas/liquid<br />

reactor<br />

Jet-impingement<br />

reactor<br />

Rotating packed-bed<br />

reactor<br />

Static mixers<br />

Compact heat<br />

exchangers<br />

Microchannel heat<br />

exchangers<br />

Rotor/stator mixers<br />

Rotating packed beds<br />

Centrifugal adsorber<br />

Process intensification<br />

Equipment Methods<br />

Multifunctional<br />

reactors<br />

Examples<br />

Reverse-flow<br />

reactors<br />

Reactive distillation<br />

Reactive extraction<br />

Reactive crystallisation<br />

Chromatographic<br />

reactors<br />

Periodic separating<br />

reactors<br />

Membrane reactors<br />

Reactive extrusion<br />

Reactive comminution<br />

Fuel cells<br />

Fig. 2.14 Adapted from Ref. 37. (Reproduced with permission<br />

of American Institute of Chemical Engineers.)<br />

example where process intensification has led to<br />

significant improvements in throughput overcoming<br />

the need for additional plant to be built. Formation<br />

of a gel by-product necessitated frequent shut down<br />

of a Dow Corning process involving a gas/liquid reaction<br />

in a packed column [38]. Owing to increased<br />

demand, the company were faced with building<br />

another plant at a cost of some US$5 million or<br />

making significant improvements in productivity.<br />

The key problems were identified as poor gas–liquid<br />

mixing and mass transfer, which were readily solved<br />

at a cost of US$22000 by adding a static mixer.<br />

The result was a 42% increase in productivity, avoid-<br />

72<br />

Hybrid<br />

separations<br />

Membrane absorption<br />

Membrane distillation<br />

Adsorptive distillation<br />

Alternative<br />

energy<br />

sources<br />

Centrifugal fields<br />

Ultrasound<br />

Solar energy<br />

Microwaves<br />

Electric fields<br />

Plasma technology<br />

Dynamic<br />

ing the requirement for a new plant together with<br />

all the materials and energy waste that this would<br />

entail.<br />

5 Reduction of Energy Requirement<br />

Other<br />

methods<br />

More than 75% of the world’s energy comes from<br />

fossil resouces [39], with approximately half of the<br />

rest coming from biomass and the remainder coming<br />

from non-carbon sources. This dependency on fossil<br />

fuels has two major consequences:<br />

(1) It is leading to rapidly diminishing reserves of<br />

this valuable non-renewable resource.<br />

(2) It is contributing to the increasing concentration<br />

of CO2 in the atmosphere.


Although these consequences are not in dispute,<br />

there is some dispute over the significance of both<br />

rising CO2 levels and the level of contribution made<br />

by the burning of fossil fuels [40]. Governments of<br />

most countries now accept that reduction of both the<br />

use of fossil fuels and CO2 emissions will be of environmental<br />

benefit and agreements and legislation<br />

are being put in place to meet these objectives<br />

[41].<br />

Energy requirements of chemical reactions frequently<br />

are overlooked at the R&D stage and, for all<br />

but the largest commodity processes, were not considered<br />

seriously at the production stage either, at<br />

least until the oil crisis in the 1970s. As energy has<br />

become more expensive and legislative drivers have<br />

encouraged greater energy efficiency and conservation,<br />

we have seen significant changes in process<br />

design. Many of these changes have focused on engineering<br />

aspects, such as using hot process streams<br />

from one part of a process to heat up incoming raw<br />

materials, whereas a combination of process reengineering<br />

and catalysis has led to energy savings<br />

in many large-scale chemical processes. Even so,<br />

energy conservation is one of the most ignored of the<br />

12 Principles of Green Chemistry, especially by<br />

chemists!<br />

5.1 Some energy efficiency improvements<br />

Ammonia has been synthesised chemically for<br />

almost 100 years. The original electric ark process<br />

operated at temperatures of over 3000°C and was<br />

highly inefficient. The Haber process was a huge leap<br />

for energy efficiency, brought about by the use of<br />

a reduced magnetite catalyst [42]. Although the<br />

underlying principles of the Haber process have<br />

changed little, the energy consumption of the<br />

process is now less than 40% of the original process<br />

[43].<br />

Initially the energy utilisation of the process was<br />

less than 20%, however with the replacement of coal<br />

by oil, and later gas, as the preferred feedstock the<br />

energy efficiency rose to around 60%. Optimisation<br />

of turbine equipment, the steam distribution networks<br />

and the design of radical flow converters<br />

with small-sized catalyst particles have made significant<br />

contributions to energy efficiency improvements<br />

without significant changes being made to the<br />

chemistry.<br />

Principles of Sustainable and Green Chemistry 21<br />

73<br />

The production of sulfuric acid has gone through<br />

similar historical improvements [43]. The main<br />

energy savings have been made in the production of<br />

sulfur dioxide, which is the initial step in the process.<br />

Originally this was produced by roasting the ore<br />

(pyrites) in multiple hearth furnaces and later rotary<br />

kilns, the energy produced being lost to the surroundings.<br />

The development of fluidised bed technology<br />

enabled more than 50% of the excess energy<br />

to be recovered and used to raise steam. Many<br />

modern plants use sulfur (recovered from oil and<br />

gas) as the feedstock and this produces much cleaner<br />

SO2, eliminating the requirement for a cleaning step<br />

and saving further energy.<br />

When considering the eco-efficiency and competitiveness<br />

of competing processes it is vital that the<br />

energy requirements of the process are considered.<br />

Unfortunately this detailed information is not readily<br />

available for most small- to medium-scale processes.<br />

As a striking example of how the energy requirements<br />

for producing a given chemical can vary from<br />

process to process, let us consider titanium dioxide<br />

production. The annual production of TiO2 is<br />

approximately 4.5 million tonnes, made via two<br />

competing processes—the sulfate process and the<br />

chloride process.<br />

The sulfate process essentially involves three<br />

stages:<br />

(1) Dissolution of the ore (ilmenite) in sulfuric acid<br />

and removal of iron impurities<br />

(2) Formation of hydrated TiO2 by treatment of the<br />

sulfate with base<br />

(3) Dehydration in a calciner<br />

All three stages are energy intensive [44].<br />

By contrast, the chloride process can, for simplicity,<br />

be broken down into two steps:<br />

(1) Chlorination of the ore with Cl2 and purification<br />

of TiCl4 by distillation<br />

(2) Oxidation by burning<br />

Overall, the chloride process is much less energy<br />

intensive (by a factor of around 5 [44]), one of the<br />

main reasons being the avoidance of large amounts<br />

of water that need to be removed by energyintensive<br />

evaporation.<br />

With such a huge energy differential it could be<br />

assumed that the chloride process should have shutdown<br />

economics, however around 40% of TiO2 is


22 Chapter 2<br />

still produced via the sulfate route, although this is<br />

diminishing. There are two main reasons for this:<br />

the sulfate process can use lower grade and therefore<br />

less-expensive ores; and it produces anatase pigments<br />

as well as rutile, which is the sole product of<br />

the chloride process.<br />

5.2 Alternative energy sources<br />

The energy required to bring about chemical reactions<br />

is supplied largely by external thermal sources<br />

of heat, such as steam, hot oil and electrical heating<br />

elements. When designing a process for energy efficiency<br />

these conventional energy sources, which do<br />

not target the energy, may not be the most efficient<br />

and alternatives should be considered. There is<br />

currently growing interest in alternative sources of<br />

energy that can target specific molecules or bonds,<br />

giving both energy savings and improved selectivity.<br />

Such alternative energy sources include microwaves<br />

and photochemical, ultrasonic and electrochemical<br />

sources, some of which are discussed in detail in<br />

other chapters of this book. Industrial manufacturing<br />

processes using electrochemistry [45] (the<br />

obvious example being chlorine/sodium hydroxide<br />

manufacture) and, to a somewhat lesser extent, photochemistry<br />

[46] (e.g. the synthesis of vitamin D3, as<br />

discussed in most photochemistry textbooks) have<br />

been used for many years, with a great deal of<br />

success for niche products. Others, such as the use<br />

of microwave reactors, are still confined largely<br />

to the R&D laboratory. One fairly rare example of<br />

microwave energy being used for chemical production<br />

is in the vulcanisation of rubber [47], where<br />

heat-up rates can be up to 100 times faster than<br />

when conventional heating methods are used. As<br />

well as saving energy, process productivity is greatly<br />

improved and the rubber obtained is less contaminated<br />

than that produced using a liquid curing<br />

medium.<br />

Although strictly speaking outside the scope of<br />

green chemistry, the importance of photochemical<br />

and electrochemical techniques to remediation and<br />

74<br />

end-of-pipe technologies should not be overlooked.<br />

For example, photocatalysis plays an important role<br />

in the purification and treatment of wastewater [48],<br />

whereas the use of electrochemical techniques for<br />

the recovery of heavy metals from electroplating<br />

processes is becoming widespread [49].<br />

6 Reduction of Risk and Hazard<br />

6.1 Inherently safe design<br />

December 1984 saw the worlds’ worst chemical disaster,<br />

with over 3000 people killed and 50000 people<br />

injured. The name Bhopal became synonymous with<br />

all that was bad about the chemical industry [50]<br />

and the repercussions from the tragedy transformed<br />

industries’ views on risk and hazard forever. The<br />

product made at Bhopal was the insecticide carbaryl<br />

(Fig. 2.15); although the chemistry involved was<br />

fairly simple, it involved the use of two highly<br />

hazardous chemicals, phosgene and methylisocyanate<br />

(MIC).<br />

The immediate cause of the accident was the<br />

ingress (during routine maintenance) of a large<br />

quantity of water into a storage tank containing up<br />

to 60 tonnes of intermediate MIC. This caused a large<br />

increase in temperature and pressure, eventually<br />

causing the storage tank to explode and a toxic gas<br />

cloud containing MIC and its hydrolysis products,<br />

including hydrogen cyanide, to be released over the<br />

nearby town.<br />

It is very easy to blame people, procedures and<br />

equipment failure in cases such as this but statistically<br />

these will always occur because people are<br />

subject to human error, procedures can always be<br />

improved with hindsight and even the most wellengineered<br />

equipment will fail eventually. It is much<br />

more beneficial to identify the real root cause of the<br />

problem and to eliminate it. In this case the root<br />

cause was the large-scale storage of toxic MIC, so the<br />

question is: could carbaryl have been manufactured<br />

efficiently without MIC being stored? The answer<br />

is yes, but that is irrelevant; what is relevant is that<br />

Fig. 2.15 The Bhopal route to carbaryl.


such questions are addressed seriously at the process<br />

design stage—the concept of ‘inherently safer design’<br />

[51].<br />

In simple terms, the risk to the environment,<br />

human life, etc. is a function of the hazard and the<br />

exposure to that hazard. Conventionally, for a given<br />

known hazard (e.g. the toxic effects of MIC) the risk<br />

has been controlled by reducing exposure through<br />

protective equipment, safety devices and other<br />

control methods. The basic concept of inherently<br />

safer design is that instead of controlling exposure it<br />

is the hazards that are, as far as possible, designed<br />

out of the process—the ‘what you don’t have can’t<br />

harm you approach’. In the Bhopal case such a<br />

design would have avoided the storage of both MIC<br />

and phosgene. Phosgene would have been made on<br />

site (the ‘just in time’ concept) and the use of MIC<br />

could have been avoided altogether by initially reacting<br />

naphthol with phosgene and subsequently reacting<br />

the so-formed chloroformate with methylamine.<br />

Adipic acid manufacture provides a suitable case<br />

study that not only involves inherently safe design<br />

but also embraces many of the principles of green<br />

chemistry discussed above.<br />

World production of adipic acid is approximately<br />

2 milliontyear -1 , the majority of which goes into the<br />

manufacture of Nylon 6:6 [52]. More than 90%<br />

of adipic acid is manufactured by the oxidation<br />

of cyclohexane via a two-stage process (Fig. 2.16)<br />

involving initial oxidation with air using a cobalt<br />

naphthenate or boric acid catalyst followed by oxidation<br />

of the ketone/alcohol mix with an excess of<br />

nitric acid [53].<br />

Although this process has been operated successfully<br />

for many years by many large chemical<br />

companies, there are many associated green and sustainable<br />

issues connected with it:<br />

(1) The Flixborough disaster in 1974 occurred in a<br />

cyclohexane oxidation plant operating at 150°C<br />

and 10bar pressure. In order to get high selec-<br />

Fig. 2.16 Conventional route to adipic<br />

acid.<br />

Principles of Sustainable and Green Chemistry 23<br />

75<br />

tivity, the plant needed to operate at low<br />

conversions (


24 Chapter 2<br />

encouraging our young chemists and engineers to<br />

adopt.<br />

• Is the product (adipic acid) required or is there an<br />

alternative?<br />

• Are there any alternative routes that can be<br />

evaluated and compared?<br />

• Can the ketone/alcohol mix be produced avoiding<br />

the inherent hazards involved with oxidation?<br />

• Can the ketone/alcohol mix be produced with<br />

lower risk?<br />

• Can the large hydrocarbon inventory be reduced?<br />

• Is there an alternative to the nitric acid oxidation<br />

step?<br />

• Can nitrous oxide be recycled in the process to a<br />

useful material?<br />

• Can nitrous oxide emissions be avoided?<br />

These types of questions need to be addressed from<br />

both the technical and economic viewpoint, with the<br />

aim of identifying a more environmentally friendly<br />

product or route that is also commercially viable.<br />

Of course many of these questions have been<br />

addressed, sometimes in an attempt to improve<br />

process and eco-efficiency, and at other times industry<br />

has been forced to consider alternatives based on<br />

public and legislative pressure following Flixborough<br />

and increased concern over the greenhouse effect for<br />

example.<br />

Following pressure to reduce nitrous oxide emissions,<br />

all major adipic acid manufacturers agreed<br />

to adopt some form of nitrous oxide abatement<br />

measure by 1998. Most of these procedures involve<br />

end-of-pipe technology that overcomes the immediate<br />

problem, at some significant cost, but does not<br />

address the real issue of avoiding nitrous oxide production.<br />

Some of the abatement technologies involved<br />

are: catalytic reduction in the presence of<br />

methane to nitrogen and carbon dioxide; catalytic<br />

dissociation into nitrogen and oxygen; and oxidation<br />

to nitric oxide, which can be used to make nitric acid.<br />

76<br />

In the cyclohexane oxidation stage, small<br />

improvements to rate and conversion have been<br />

made by improved design of air/liquid mixers and<br />

heat exchangers (process intensification) [51], but<br />

the inherent hazards associated with stage 1 of the<br />

process largely remain.<br />

Perhaps the preferred answer to achieving a<br />

greener, more inherently safe process is to find an<br />

alternative route. Production of cyclohexanol by catalytic<br />

hydrogenation of phenol proceeds in high<br />

yield and high selectivity, and at first sight overcomes<br />

the cyclohexane oxidation issues. However, the<br />

starting material for both routes is benzene, and<br />

phenol production involves a similar oxidation step<br />

(see cumene route discussed above). Thus, it may<br />

be argued that cyclohexanol production has been<br />

‘greened’ but the issues have been put back a stage.<br />

On the other hand, it could be argued that because<br />

phenol is produced anyway the number of oxidation<br />

plants required has been reduced and therefore the<br />

overall risk has been reduced. Although the phenol<br />

route is practised, unfavourable economics have prevented<br />

widespread adoption of this process.<br />

What could be—at least if the economics were<br />

more favourable—the perfect green and sustainable<br />

solution has been identified by Frost [56]. The<br />

process involves synthesis of adipic acid from glucose<br />

via catechol (Fig. 2.17) using a genetically modified<br />

Escherichia coli biocatalyst.<br />

Several other processes, with varying degrees of<br />

greenness, have been developed over the years,<br />

including carboalkoxylation of butadiene [57] and<br />

dicarbonylation of 1,4-dimethoxybut-2-ene [58],<br />

but for various reasons these have not been commercially<br />

successful.<br />

The most successful move towards sustainability<br />

has been made by Asahi Kasei Corp. [59], who have<br />

developed a commercial route to cyclohexanol based<br />

on the hydration of cyclohexene (Fig. 2.18). The<br />

process relies on the use of a novel high-silica H-<br />

Fig. 2.17 Green route to adipic acid.


Fig. 2.18 Asahi route to cyclohexanol.<br />

ZMS-5 catalyst and the hydrophobic properties of<br />

this catalyst enable selective adsorption of cyclohexene<br />

to take place.<br />

6.2 Alternative solvents<br />

The use of benign solvents is of vital importance to<br />

the development of green chemistry. The concept is<br />

introduced here for convenience but could have<br />

been discussed equally under waste minimisation<br />

and materials reduction. The use of organic solvents<br />

in industry is widespread and, although they play a<br />

valuable ‘enabling’ role, they are responsible for significant<br />

amounts of pollution to air and water. To<br />

control the pollution, industry is spending increasing<br />

amounts on volatile organic compounds and other<br />

effluent control measures. Many organic solvents are<br />

also toxic, necessitating the use of personal protective<br />

equipment; although these are being phased<br />

out by industry, many are still commonly used in<br />

research.<br />

The greenest solution is to use no solvent at all, a<br />

concept not considered by most practising chemists,<br />

at least where solid reactants are concerned. Raston<br />

[60] has reported a striking example of what can<br />

be achieved. The widely used aldol condensation<br />

usually is carried out in solvents such as ethanol,<br />

however Raston’s group have demonstrated that<br />

selective, high-yielding conversions could be<br />

achieved by grinding together a solid carbonyl compound<br />

with sodium hydroxide using a pestle and<br />

mortar.<br />

Water is also a much under-utilised solvent for<br />

organic reactions and offers the potential for cheaper<br />

and safer processes. Water is especially worth considering<br />

as a solvent for high-temperature reactions<br />

where the ionic product of water is high, making it<br />

a stronger acid and base, and where the polarity is<br />

lower, making it a better solvent for organic com-<br />

Principles of Sustainable and Green Chemistry 25<br />

77<br />

pounds. Largely because of legislation on volatile<br />

organic compounds the use of water as a solvent<br />

is being more widely studied by industry, where<br />

the cost and environmental benefits are becoming<br />

realised [61].<br />

Supercritical carbon dioxide, ionic liquids and fluorous<br />

biphase systems all are being studied actively<br />

as eco-friendly alternatives to the use of organic solvents.<br />

These solvents are discussed elsewhere; all<br />

have potential benefits and some disadvantages that<br />

should be considered.<br />

Having moved away from highly toxic solvents<br />

such as benzene and carbon tetrachloride some years<br />

ago, we are now entering a new era where there<br />

are real alternatives being developed to common<br />

solvents such as toluene, dichloromethane and<br />

dimethylformamide. When developing a process, all<br />

these alternatives should be considered.<br />

Approaches such as life-cycle assessment are vital<br />

in enabling sound decisions to be made. For<br />

example, when considering the use of supercritical<br />

carbon dioxide against toluene, one should take<br />

account of the energy involved in generating<br />

and recirculating the supercritical CO2 as well as<br />

assessing the potential pollution risk involved with<br />

using toluene and the comparative hazards of<br />

flammability versus the use of high-pressure equipment,<br />

etc.<br />

7 Conclusions<br />

Chemicals, subtly disguised as medicines, pesticides,<br />

cosmetics, clothing, fuel, protective packaging,<br />

household items, etc., have contributed enormously<br />

to the improved quality of life and increased<br />

longevity the human race has experienced during<br />

the twentieth century. Unfortunately, without intent<br />

but sometimes without adequate knowledge or<br />

thought, chemists and engineers, through the products<br />

and processes devised, have caused significant<br />

harm to the environment.<br />

The challenge to chemists and engineers in the<br />

twenty-first century is to continue to reap all the<br />

social and economic benefits that chemistry has to<br />

offer but without causing damage to the environment<br />

(in the widest sense) or preventing our children<br />

and grandchildren from doing the same in the<br />

next century. This is what green and sustainable<br />

chemistry is all about, but how can we achieve this<br />

worthy goal? The answer is complex, involves all of


26 Chapter 2<br />

society and will involve aspects of all of the following,<br />

coming together as a coherent package:<br />

(1) Technology. A significant amount of valuable<br />

green technology exists, as is evident from publications<br />

such as this and the references therein.<br />

Further green technology needs to be developed<br />

and, more importantly, this technology needs<br />

to be commercially viable. In the short- to<br />

medium-term future we are likely to see significant<br />

reductions in the use of hazardous volatile<br />

organic solvents through increased use of benign<br />

alternatives. More widespread use of catalysts<br />

to reduce waste and improve selectivity, particularly<br />

in the speciality and pharmaceutical<br />

industry, undoubtedly will be forthcoming.<br />

Engineeringwise, we will see something of a<br />

design revolution, reducing materials and<br />

energy use and improving safety. However, in<br />

terms of sustainability we require significant<br />

advances in the development of renewable<br />

sources of energy.<br />

(2) Education. Increased environmental awareness<br />

and the role that science can play need to be<br />

taught at an early age. In tertiary education the<br />

principles of green chemistry need to underpin<br />

the whole undergraduate and graduate course<br />

structure and not be taught as a separate course.<br />

It is the culture and mind-set that we teach students<br />

of today that will determine the greenness<br />

of tomorrow’s processes and products.<br />

(3) Society in general has a huge role to play. Consumer<br />

power is a major force in determining the<br />

products that industry produces. Sometimes sustainability<br />

comes at a price, either economic- or<br />

performancewise; it is the informed consumer<br />

who needs to make this choice, e.g. water-based<br />

versus solvent-based paint. As stakeholders in<br />

industry, individuals and organisations also can<br />

have an influential role in ensuring that sustainable<br />

practices are adopted.<br />

(4) Governments have a role to play in formulating<br />

policies to encourage sustainable development<br />

and in ensuring that there is an even competitive<br />

playing field throughout the world.<br />

Governments also should be encouraging the<br />

development of green technologies through R &<br />

E funding mechanisms.<br />

(5) Industry has the prime purpose of adding value<br />

to shareholder investment. However, there is the<br />

78<br />

growing realisation that the key to future success<br />

lies with not one but three bottom lines: economic,<br />

environmental and social. The costs<br />

of causing environmental harm are becoming<br />

unacceptably high in every sense. This, coupled<br />

with increasing knowledge of the long-term<br />

effects that chemicals have on the environment<br />

and the technical advances are helping to ensure<br />

that new products and processes are much<br />

greener than those being replaced.<br />

The goal of a green and sustainable society cannot<br />

be achieved overnight; the path is long and uncertain.<br />

Undoubtedly we will take a few wrong turns<br />

along the way, but there is growing agreement on<br />

the general way forward. Although science can<br />

help move towards the ideal of completely avoiding<br />

risk, hazard, pollution, etc., society at large will<br />

ultimately determine what is acceptable in terms<br />

of the cost–benefit issues posed by sustainable<br />

development.<br />

Although, as discussed at the beginning of the<br />

chapter, the modern ideas of sustainable and green<br />

chemistry are only around ten years old, we should<br />

not lose sight of the fact that some of the objectives<br />

have been pursued for many years. A quote from<br />

R. W. Hofmann, the first President of The Royal<br />

College of Chemistry, London, made in 1848 is just<br />

as relevant today as it was then:<br />

‘In an ideal chemical factory there is, strictly<br />

speaking, no waste but only product. The better a<br />

real factory makes use of its waste, the closer it<br />

gets to its ideal, the bigger is the profit.’<br />

References<br />

1. World Commission on Environment and Development.<br />

Our Common Future. Oxford University Press,<br />

Oxford, 1987.<br />

2. Basiago, A. D. Sustain. Dev., 1995, 3, 109.<br />

3. Anastas, P. T., & Warner, J. C., Green Chemistry, Theory<br />

and Practice. Oxford University Press, Oxford, 1998.<br />

4. Environmental Protection Expenditure by UK Industry.<br />

Survey of 1997 Expenditure, final report to the DETR.<br />

ECOTEC Research and Consulting, Birmingham, UK,<br />

1999.<br />

5. Lancaster, M. Environ. Bus. Mag., 1999, May, 34.<br />

6. MORI. The Public Image of the Chemical Industry, report<br />

for the CIA. MORI, London, 1999.<br />

7. Trost, B. M. Science, 1991, 254, 1471.<br />

8. Sheldon, R. A. Chemtech, 1994, March, 38.<br />

9. Sheldon, R. A. J. Chem. Tech. Biotechnol., 1997, 68, 381.


10. Lancaster, M. Green Chem., 2000, 2, G65.<br />

11. For a full introduction to pericyclic reactions, see:<br />

Fleming, I. Frontier Orbitals and Organic Chemical Reactions.<br />

J. W. Arrowsmith, Bristol, 1992.<br />

12. Heaton, A. The Chemical Industry, Chapt. 7. Blackie<br />

Academic & Professional, Glasgow, 1996.<br />

13. Saksena, A. K., Girijavallabhan, V. M., Chen, Y. T., Jao,<br />

E., et al. Heterocycles, 1993, 35, 129.<br />

14. Kahn, P. H., & Cossy, J. Tetrahedron Lett., 1999, 40,<br />

8113.<br />

15. Lavallee, J. F., & Deslongchamps, P. Tetrahedron Lett.,<br />

1988, 29, 6033.<br />

16. German patent DE 4 326 952 to BASF, 1993.<br />

17. German patent DE 4 326 953 to BASF, 1993.<br />

18. Wang, Z., Campagna, S., Xu, G., Pierce, M. E., Fortunak,<br />

J. M., & Confalone, P. N. Tetrahedron Lett., 2000,<br />

41, 4007.<br />

19. Doyle, M. P., & Mungwall, W. S. Experimental Organic<br />

Chemistry. John Wiley, Chichester, 1980.<br />

20. European Patent Application 0284310 to Hoechst<br />

Celanese, 1988.<br />

21. Tanabe, K., & Holderich, W. F. Appl. Catal. A: General,<br />

1999, 181, 399.<br />

22. WO Patent 96/35655 to Rhodia, 1996.<br />

23. Smith, K., Musson, A., & DeBoos, G. A. Chem.<br />

Commun., 1996, 469.<br />

24. Barrett, A. G. M., Waller, F. J., Braddock, D. C., &<br />

Ramprasad, D. Chem. Commun., 1997, 613.<br />

25. Nojiri, N., & Misono, M. Appl. Catal. A, 1993, 93, 103.<br />

26. Verweij, J., & de Vroom, E. Rec. Trav. Chim. Pays-Bas,<br />

1993, 112, 66.<br />

27. Campbell, C. J., & Laherrere, J. H. Sci. Am., 1998,<br />

March, 60.<br />

28. Shay, E. G. Biomass Bioen., 1993, 4, 227.<br />

29. Ma, F., & Hanna, M. A. Bioresource Technol., 1999,<br />

70, 1.<br />

30. Adams, C., Peters, J. F., Rand, M. C., Schroer, B. J., &<br />

Pzremke, M. C. J. Am. Oil Chem. Soc., 1983, 60, 1574.<br />

31. Jackson, M. A., & King, J. W. J. Am. Oil Chem. Soc.,<br />

1996, 62, 815.<br />

32. Wasserbauer, R., Beranova, M., Vancurova, D., &<br />

Dolezel, B. Biomaterials, 1990, 11, 36.<br />

33. Environmental Data Services. ENDS Report 2000, Vol.<br />

300. ENDS, London, pp. 19–21.<br />

34. Dartee, M. Green-Tech 2000, Utrecht, April 2000.<br />

35. Okkerse, C., & van Bekkum, H. Green Chem., 1999, 1,<br />

107.<br />

36. Ramshaw, C. Green Chem., 1999, 1, G15.<br />

Principles of Sustainable and Green Chemistry 27<br />

79<br />

37. Stankiewick, A. I., & Moulijn, J. A. Chem. Eng. Prog.,<br />

2000, Jan., 23.<br />

38. Green, A., Johnson, B., & John, A. Chem. Eng., 1999,<br />

Dec., 66.<br />

39. Victor, D. Nature, 1998, 395, 837.<br />

40. Armor, J. N. J. Appl. Catal. A: General, 2000, 194/195,<br />

3.<br />

41. For a summary of the Kyoto agreement, see:<br />

www.unfccc.de<br />

42. Adams, C. Chem. Ind., 1999, 19, 740.<br />

43. Muller, H. In Ullmans Encyclopedia of Industrial Chemistry,<br />

5th edn, Vol. 25. VCH, Wurzburg, 1994, pp.<br />

635–703.<br />

44. Brown, H. L., Hamel, B. B., & Hedman, B. A. Energy<br />

Analysis of 108 Industrial Processes. Fairmont Press,<br />

Lilburn, 9A, 1996.<br />

45. Scott, K. Electrochemical Processes for Clean Technology.<br />

Royal Society of Chemistry, London, 1995.<br />

46. Bouchy, A., Andre, J. C., George, E., & Viriot, M. L. J.<br />

Photochem. Photobiol., 1989, 48, 447.<br />

47. Meredith, R. J. J. Elast. Plast., 1996, 8, 191.<br />

48. Ollis, D. F., & Al-Ekabi, H. Photocatalytic Purification and<br />

Treatment of Water and Air, Trace Metals in the Environment,<br />

Vol. 3. Elsevier, Amsterdam, 1993.<br />

49. Robinson, D., & Walsh, F. C. Hydrometallurgy, 1991,<br />

26(93), 115.<br />

50. Everest, L. Behind the Poison Cloud: Union Carbide’s<br />

Bhopal Massacre. Banner Press, Birmingham, AL,<br />

1986.<br />

51. Kletz, T. Process Plants. A Handbook for Inherently Safer<br />

Design. Taylor & Francis, London, 1998.<br />

52. Eur. Chem. News, 1999, 71, 18.<br />

53. Davis, D. D. In Ullmans Encyclopedia of Industrial Chemistry,<br />

5th edn, Vol. A1. VCH, New York, 1985, pp.<br />

269–278.<br />

54. Scott, R. Process Optimisation, Symposium Series 100.<br />

Institution of Chemical Engineers, Rugby, UK, 1987.<br />

55. Thiemens, M. H., & Trogler, W. C. Science, 1991, 251,<br />

932.<br />

56. Draths, K. M., & Frost, J. W. In Green Chemistry, Frontiers<br />

in Benign Chemical Synthesis and Processes (Anastas,<br />

P. T., & Williamson, T. C. eds). Oxford University Press,<br />

Oxford, 1998.<br />

57. US Patent 4 259 520 to BASF, 1981.<br />

58. Chem. Eng. News, 1984, 62(18), 28.<br />

59. DE Patent 3441072 to Asahi Kasei Corp., 1985.<br />

60. Raston, C. L., & Scott J. L., Green Chem., 2000, 2, 49.<br />

61. Cook, S. J. Green Chem., 1999, 1, 138.


BYLAAG 3<br />

Baird, C and Cann, M, 2008, Environmental chemistry, 4 th ed., W. H. Freeman and Company: New York,<br />

NY, p. 601 – 660.<br />

80


C H A P T E R 14<br />

THE POLLUTION AND<br />

PURIFICATION OF WATER<br />

In this chapter, the following introductory chemistry<br />

topics are used:<br />

Acid–base and equilibrium concepts and calculations; pH<br />

Basic structural organic chemistry (as in the Appendix)<br />

Oxidation numbers; redox half-reactions<br />

Catalysis<br />

Distillation<br />

Background from previous chapters used in this chapter:<br />

Maximum contaminant levels (Chapter 10)<br />

BOD (Chapter 13)<br />

VOCs (Chapter 3)<br />

Adsorption (Chapter 4)<br />

Photochemical reactions; UV light (Chapters 1–5)<br />

Free radicals (Chapter 1)<br />

BTX hydrocarbons (Chapter 7)<br />

ppm concentration scale in water (Chapter 10)<br />

No effects level, NOEL (Chapter 11)<br />

Introduction<br />

The pollution of natural waters by both biological and chemical contaminants<br />

is a worldwide problem. There are few populated areas, whether in developed<br />

or undeveloped countries, that do not suffer from one form of water pollution<br />

or another. In this chapter, we shall survey the various methods—both traditional<br />

and innovative—by which water can be purified. We begin by discussing<br />

techniques that are used to purify drinking water from relatively<br />

uncontaminated sources, and then consider the pollution and remediation of<br />

81<br />

601


602 Chapter 14 The Pollution and Purification of Water<br />

groundwater and of sewage and wastewater. Lastly, we investigate modern<br />

advanced techniques whereby polluted air and water can be cleansed.<br />

Water Disinfection<br />

The quality of “raw” (untreated) water, whether drawn from surface water or<br />

groundwater, that is intended eventually for drinking varies widely, from<br />

almost pristine to highly polluted. Because both the type and quantity of pollutants<br />

in raw water vary, the processes used in purification also vary from<br />

place to place. The most commonly used procedures are shown in schematic<br />

form in Figure 14-1. Before discussing the major topic of disinfection, we shall<br />

discuss the various nondisinfection steps that are often taken in the overall<br />

purification process.<br />

Aeration of Water<br />

Aeration is commonly used in the improvement of water quality. Municipalities<br />

aerate drinking water that is drawn from underground aquifers in order<br />

to remove dissolved gases such as the foul-smelling hydrogen sulfide, H2S, and organosulfur compounds, as well as volatile organic compounds, some of<br />

which may have a detectable odor. Aeration of drinking water also results in<br />

reactions that produce CO2 from the most easily oxidized organic material. If<br />

necessary for reasons of odor, taste, or health, most of the remaining organics<br />

can be removed by subsequently passing the water over activated carbon,<br />

although this process is relatively expensive, so rather few communities use it<br />

(see Box 14-1). Another advantage to aeration is that the increased oxygen<br />

content of water oxidizes water-soluble Fe 2 to Fe 3 , which then forms insoluble<br />

hydroxides (and related species) that can be removed as solids.<br />

Fe 3 3 OH 9: Fe(OH) 3(s)<br />

(Recall that ions listed in equations without a state specified are assumed to<br />

be in aqueous solution.)<br />

After aeration, colloidal particles in the water are removed. If the water<br />

is excessively hard, calcium and magnesium are removed from it before the<br />

final stages of disinfection and the addition of fluoride. All these procedures<br />

are described below (see Figure 14-1).<br />

Removal of Calcium and Magnesium<br />

If the water comes from wells in areas having limestone bedrock, it will contain<br />

significant levels of Ca 2 and Mg 2 ions, which are usually removed during<br />

processing since these ions can interfere with soaps and detergents used<br />

by consumers for washing. Calcium can be removed from water by addition of<br />

phosphate ion in a process analogous to that discussed later for phosphate<br />

removal; here, however, phosphate is added in order to precipitate the<br />

82


BOX 14-1 Activated Carbon<br />

Activated carbon (activated charcoal) is a<br />

very useful solid for purifying water of<br />

small organic molecules present in low concentrations.<br />

The ability of this material to<br />

remove contaminants from water and to<br />

improve its taste, color, and odor has been<br />

known for a long time; indeed, the ancient<br />

Egyptians used charcoal-lined vessels to store<br />

water for drinking purposes.<br />

Activated carbon is produced by anaerobically<br />

charring a high-carbon-content material<br />

such as peat, wood, or lignite (a soft brown<br />

coal) at temperatures below 600°C, followed by<br />

a partial oxidation process using carbon dioxide<br />

or steam at a slightly higher temperature.<br />

The removal of contaminants by activated<br />

carbon is a physical adsorption process and<br />

therefore is reversible if sufficient energy is<br />

applied. The characteristic that makes activated<br />

carbon such an excellent adsorber is its<br />

huge surface area, about 1400 m 2 /g. This surface<br />

is internal to the individual carbon particles,<br />

so that crushing the material neither<br />

Aeration Settling and<br />

precipitation<br />

Air<br />

Al or Fe salt to<br />

precipitate colloids<br />

Hardness<br />

removal<br />

Phosphate<br />

FIGURE 14-1 The common stages of purification of drinking water.<br />

Water Disinfection 603<br />

Disinfection<br />

Ammonia<br />

and fluoride<br />

Ca 2+ pptd. as phosphate Cl 2 or ozone or ClO 2<br />

83<br />

increases nor decreases the area. The internal<br />

structure of the solid involves series of channels<br />

(pores) of progressively decreasing size<br />

that are produced by the charring and partial<br />

oxidation processes. The internal sites where<br />

adsorption occurs are large enough only for<br />

small molecules, including chlorinated solvents.<br />

At the typical ppm concentrations<br />

found for organic contaminants in water, each<br />

gram of activated carbon can adsorb a few<br />

percent of its mass in contaminants such as<br />

chloroform and the dichloroethenes as well as<br />

much higher masses of TCE, PCE, and pesticides<br />

such as dieldrin, heptachlor, and DDT.<br />

Once a sample of activated carbon has<br />

reached near-saturation in terms of adsorbed<br />

organics, three alternatives are available. It<br />

can be simply disposed of in a landfill, it can be<br />

incinerated to destroy it and the adsorbed contaminants,<br />

or it can be heated to rejuvenate the<br />

surface by driving off the organic pollutants,<br />

which can then be incinerated or catalytically<br />

oxidized.<br />

Consumer<br />

Water<br />

Suspended<br />

particles


604 Chapter 14 The Pollution and Purification of Water<br />

calcium ion. More commonly, calcium ion is removed by precipitation and<br />

filtering of the insoluble salt calcium carbonate, CaCO 3 . The carbonate<br />

ion is either added as sodium carbonate, Na 2 CO 3 , or if sufficient HCO 3 is<br />

naturally present in the water, hydroxide ion, OH , is added in order to<br />

convert dissolved bicarbonate ion to carbonate:<br />

OH HCO 3 9: CO3 2 H2 O<br />

Ca 2 CO 3 2 Δ CaCO3 (s)<br />

Magnesium ion precipitates as magnesium hydroxide, Mg(OH) 2, when the<br />

water is made sufficiently alkaline, i.e., when the OH ion content is increased.<br />

After removal by filtration of the solid CaCO 3 and Mg(OH) 2 , the pH of the<br />

water is readjusted to near-neutrality by bubbling carbon dioxide into it.<br />

PROBLEM 14-1<br />

Ironically, calcium ion is often removed from water by adding hydroxide ion<br />

in the form of Ca(OH) 2 . Deduce a balanced chemical equation for the reaction<br />

of calcium hydroxide with dissolved calcium bicarbonate, Ca(HCO 3 ) 2 ,<br />

to produce insoluble calcium carbonate. What molar ratio of Ca(OH) 2 to<br />

dissolved calcium should be added to ensure that almost all the calcium is<br />

precipitated?<br />

Disinfection to Prevent Illness<br />

In terms of causing immediate sickness and even death, biological contaminants<br />

of water are almost always much more important than chemical ones.<br />

For that reason, we begin our discussion of the purification of water by extensively<br />

discussing its disinfection, i.e., the elimination of microorganisms that<br />

can cause illness.<br />

Many of the microorganisms in raw water are present as a result of contamination<br />

by human and animal feces. The microorganisms are principally<br />

• bacteria, including those of the Salmonella genus, one species of which<br />

causes typhoid. In this category is also Escherichia coli O157:H7, whose transmission<br />

in water has caused a number of deaths in recent years, including<br />

those from an outbreak in Walkerton, Ontario, in 2000;<br />

• viruses, including polio viruses, the hepatitus-A virus, and the Norwalk<br />

virus; and<br />

• protozoans (single-celled animals), including Cryptosporidium and Giardia<br />

lamblia.<br />

Because many microorganisms of these three types are pathogenic, causing<br />

mild to serious and sometimes fatal illnesses, they must be largely removed<br />

from water before it is suitable for drinking.<br />

84


Notwithstanding well-known techniques for water disinfection, many of<br />

which have been used extensively for more than a century in developed<br />

countries, there are still about 1 billion people in the world who do not yet<br />

have access to safe drinking water. According to the World Health Organization,<br />

about 4500 children die daily from the consequences of polluted water<br />

and inadequate sanitation.<br />

Filtering of Water<br />

In addition to dissolved chemicals, the raw water that is obtained from<br />

rivers, lakes, or streams contains a multitude of tiny particles, some of which<br />

consist of or contain microorganisms. Many of the small, suspended particles<br />

consist of clay, resulting from the erosion of soil and rock, whether by natural<br />

forces or due to plowing of land for agriculture, mining, or commercial or<br />

housing development. The suspended particles increase water’s turbidity and<br />

thereby reduce the ability of light to penetrate deeply enough to support<br />

photosynthesis.<br />

The larger of the particles suspended in water are often removed by simply<br />

filtering it. Indeed, the filtration of water by passing it through a bed of<br />

sand is the oldest form of water purification known, dating back to ancient<br />

times. The sand retains suspended solids of all types, including microorganisms,<br />

down to about 10 mm in size.<br />

Recently it was realized that forcing raw water through filters having<br />

especially small openings can be used instead of chemicals or light to disinfect<br />

water of some viruses and bacteria, and even some dissolved chemicals, by<br />

just removing them.<br />

Removal of Colloidal Particles by Precipitation<br />

Most municipalities allow raw water to settle, since this permits large particles<br />

to settle out or to be readily separated. However, much of the insoluble matter<br />

originating from rocks and soil, and from the disintegration and decomposition<br />

of water-based plants and animals, will not precipitate spontaneously<br />

since it is suspended in water in the form of colloidal particles. These are particles<br />

that have diameters ranging from 0.001 to 1 mm and consist of groups of<br />

molecules or ions that are weakly bound together. These groups dissolve as a<br />

unit, rather than breaking up and dissolving as individual ions or molecules. In<br />

many cases, the individual units within a colloidal particle are spatially organized<br />

such that the surface of the particles contains ionic groups. The ionic<br />

charges on the surface of one particle repel those of like charge on neighboring<br />

particles, preventing their aggregation and subsequent precipitation.<br />

Colloidal particles must be removed from drinking water for both aesthetic<br />

and health and safety reasons. To capture the colloidal particles, a<br />

small amount of either iron(III) sulfate, Fe2(SO4) 3, or aluminum sulfate,<br />

85<br />

Water Disinfection 605


606 Chapter 14 The Pollution and Purification of Water<br />

Al 2 (SO 4 ) 3 (“alum”), is deliberately dissolved in the water. By then making<br />

the water neutral or alkaline in pH (7 and up), both the Fe 3 and Al 3 ions<br />

produced from the salts form gelatinous hydroxides that physically incorporate<br />

the colloidal particles and form a removable precipitate. The water is<br />

greatly clarified once this precipitate has been removed. Commonly, after the<br />

removal of the colloidal particles, the water is filtered through sand and/or<br />

some other granular material.<br />

Although the idealized formulas of the precipitates are Fe(OH) 3 and<br />

Al(OH) 3 , the actual situation is much more complex. For example, aluminum<br />

actually forms a polymeric cation, Al 13 O 4 (OH) 24 7 , which produces<br />

a loose network structure that is held together by hydrogen bonds. This network<br />

entraps the colloidal particles and forms the precipitate. Only if the pH<br />

rises to quite a high value does the aluminum in solution form the expected<br />

hydroxide Al(OH) 3 . Since the concentration of aluminum sulfate added to<br />

the water is only about 10 mm/L, very little residual aluminum ion is left in<br />

the treated water.<br />

PROBLEM 14-2<br />

Calculate the approximate number of atoms contained in colloidal particles<br />

of (a) 1-mm and (b) 0.01-mm diameters, assuming that their densities are similar<br />

to that of water and that the atomic mass of the atoms averages 10 g/mol.<br />

Disinfection of Water by Membrane Technology<br />

Water can be purified of most contaminant ions, molecules, and small particles,<br />

including viruses and bacteria, by passing it through a membrane in<br />

which the individual holes, called pores, are of uniform and microscopic size.<br />

The range of sizes of the various contaminants in raw water are summarized in<br />

Figure 14-2. Clearly, for a technique to be effective in providing a barrier, the<br />

pore size of the membrane must be smaller than the contaminant size.<br />

In the processes of microfiltration and ultrafiltration, a membrane or<br />

some other analogous barrier containing pores of 0.002- to 10-mm diameter<br />

(2–10,000 nm) is employed to remove larger constituents from water. The<br />

water can be forced through the barrier by pressure or can be drawn through it<br />

by suction, leaving behind the larger impurities. In one modern version of this<br />

technology, the barrier is composed of thousands of strands of plastic tubing<br />

having walls that are pierced with thousands of tiny pores of similar size.<br />

Some bacteria and colloid particles are as small as 0.1 mm and so can pass<br />

through conventional filters and even some microfilters (Figure 14-2).<br />

Viruses can be as small as 0.01 mm and therefore require at least the ultrafiltration<br />

level to eliminate them. However, filtration using membranes can be<br />

86


Pressure<br />

Microfiltration<br />

Ultrafiltration<br />

Bacteria<br />

and<br />

colloids<br />

Viruses<br />

Nanofiltration<br />

Ca 2+<br />

Mg 2+<br />

M 2+<br />

Reverse osmosis<br />

10 1 0.1 0.01 0.001 0.0001<br />

Size of contaminant ( μm)<br />

used to disinfect water if a sufficiently small pore size is used and if the water<br />

is later irradiated with ultraviolet light to eliminate any microbes that have<br />

passed through the filtration stage.<br />

Neither microfiltering nor ultrafiltering removes dissolved ions or small<br />

organic molecules. Generally speaking, before water is treated using membranes<br />

with even smaller pores (see below), it must be pretreated to remove<br />

the larger particles—especially colloids—which would otherwise foul the<br />

finer membrane by leaving deposits.<br />

Membrane systems have been developed recently that purify water of virtually<br />

all contaminants by nanofiltration. Water is pumped under pressure<br />

through fine membranes that have pores only about 1 nm wide, which therefore<br />

remove not only most bacteria and viruses, but also any larger organic<br />

molecules that would nourish the regrowth of bacteria. These nanofilters still<br />

allow water molecules to pass through the filter, since the molecules are only<br />

a few tenths of a nanometer in size. Unlike ultrafiltration, nanofiltration can<br />

be used to soften water, since hydrated divalent ions such as Ca 2 and Mg 2<br />

are larger than the pores and so do not pass through. Hydrated monovalent<br />

ions such as sodium and chloride also pass through some nanofilters, but not<br />

through ones with subnanometer pore sizes. As a consequence, some nanofilter<br />

membrane systems can be used to desalinate seawater and to help purify<br />

wastewater, as discussed later in this chapter.<br />

Na +<br />

K +<br />

87<br />

Water Disinfection 607<br />

H 2O<br />

FIGURE 14-2 Filtration of<br />

contaminants by various<br />

methods.


608 Chapter 14 The Pollution and Purification of Water<br />

Reverse Osmosis<br />

The ultimate in membrane filtration occurs in the widely used technique<br />

called reverse osmosis, sometimes called hyperfiltration. Here, water is forced<br />

under high pressure to pass through the pores in a semipermeable membrane,<br />

composed of an organic polymeric material such as cellulose acetate or triacetate<br />

or a polyamide. Since only water (and other molecules of its small size)<br />

can pass efficiently through the pores, the liquid on the other side of the<br />

membrane is purified water. The solution on the impact side of the membrane<br />

becomes more and more concentrated in contaminants as time goes on and is<br />

discarded. The procedure is called reverse osmosis because, by use of pressure,<br />

the natural phenomenon of osmosis—by which pure water would spontaneously<br />

migrate through the membrane into solution, thereby diluting it—is<br />

reversed.<br />

Particles, molecules (including small organic molecules), and ions down<br />

to less than 1 nm (0.001 mm) in size, or about 150 g/mol in mass, are removed<br />

by reverse osmosis. It is particularly useful for removing alkali and alkaline<br />

earth metal ions, as well as salts of heavy metals. Thus it is employed in hospitals<br />

and renal units to produce water that is particularly free of ions. Reverse<br />

osmosis is used on large scale for the desalination, i.e., the removal of salts,<br />

from seawater and brackish water, a topic considered in Box 14-2.<br />

BOX 14-2 The Desalination of Salty Water<br />

Desalination is the production of fresh<br />

water from salty water, often seawater, by<br />

the removal of its ions. There are more than<br />

15,000 large-scale desalination plants in operation,<br />

located in more than 125 countries.<br />

Reverse osmosis is widely used in some areas of<br />

the world, such as the Middle East, to generate<br />

drinking water from salt water.<br />

The other main commercial desalination<br />

process is the thermal distillation—evaporation—<br />

of seawater or brackish water. Desalination of<br />

seawater by evaporation is a technique that<br />

goes back to ancient times; it is especially<br />

suited even today for seawater that contains<br />

particularly high levels of dissolved salts and<br />

suspended solids in areas such as the Persian<br />

Gulf. The evaporation method is even more<br />

energy-intensive than is reverse osmosis.<br />

Modern, large-scale thermal distillation plants<br />

use energy to raise salty water to the boiling<br />

point, then reduce the air pressure above the<br />

liquid to create a partial vacuum into which<br />

the liquid readily “flash” evaporates, leaving<br />

the salt behind in the remaining liquid. The<br />

vapor is removed and condensed as desalted<br />

water. Thermal distillation plants are often<br />

incorporated within electricity-generating<br />

plants to use the low-grade waste steam from<br />

the latter as their energy source.<br />

Desalination of water is also sometimes<br />

accomplished using the technique of electrodialysis,<br />

which is described later in this chapter.<br />

88


Water destined for drinking purposes is commonly pretreated, e.g., by<br />

filtering it through sand and gravel, and passing it over activated carbon to<br />

remove the larger particles such as bacteria, etc., and treating it with chlorine,<br />

before subjecting it to reverse osmosis in order to minimize fouling and<br />

degradation of the membrane.<br />

Because of the high pressures needed to force water through the small<br />

pores in the membrane, reverse osmosis is an energy-intensive process. A<br />

pressure of about 2 atm is sufficient for portable and domestic units, but a<br />

greater force must be applied to brackish or salty water. However, advances in<br />

the engineering of large-scale desalination plants have markedly reduced<br />

energy consumption by redirecting pressure from waste brine to low-pressure<br />

incoming water.<br />

Reverse osmosis tends to be wasteful of water, since so much of it—a<br />

third to a half—is discarded. Also, the accumulated discharges of brine—<br />

sometimes called concentrate—from desalination processes of any kind can<br />

cause cumulative environmental problems, such as harming fish populations,<br />

in the immediate area of the seacoast into which it is deposited if it is not first<br />

treated. In some locations, the brine is injected into an underground saltwater<br />

aquifer. In others, the brine is left to evaporate in large outdoor pools, and<br />

the salts disposed of later.<br />

Some domestic consumers of drinking water have installed small underthe-sink<br />

reverse osmosis units to further purify their water by removing<br />

unwanted contaminants such as heavy metal cations (e.g., lead), hard water<br />

cations (calcium and magnesium), anions (e.g., nitrate and fluoride), and<br />

organic molecules from water obtained from domestic supplies. Small reverse<br />

osmosis units are also used in medical facilities for producing water that is<br />

particularly ion-free.<br />

Some bottled water has been purified and deionized by reverse osmosis,<br />

but small amounts of salts are reintroduced into it before it is sold to consumers.<br />

Drinking large amounts of deionized water is not healthy, since the<br />

ion balance in the body can be upset as a consequence.<br />

Disinfection by Ultraviolet Irradiation<br />

Ultraviolet light can also be used to disinfect and purify water. Powerful<br />

lamps containing mercury vapor whose excited atoms emit UV-C light (see<br />

Chapter 1) centered at a wavelength of 254 nm are immersed in the water<br />

flow. About 10 seconds of irradiation are usually sufficient to eliminate the<br />

toxic microorganisms, including Cryptosporidium, which is resistant to treatment<br />

by some other methods. The germicidal action of the light disrupts the<br />

DNA in microorganisms, preventing their subsequent replication and<br />

thereby inactivating the cells. At the molecular level, absorption of UV-C<br />

light results in the formation of new covalent bonds between nearby thymine<br />

units on the same strand of DNA. If sufficient such thymine dimers are<br />

89<br />

Water Disinfection 609


610 Chapter 14 The Pollution and Purification of Water<br />

formed, the DNA molecule becomes so distorted that subsequent replication<br />

of the organism is prevented.<br />

The use of ultraviolet light to purify water is complicated by the presence<br />

of dissolved iron and humic substances, both of which absorb the UV light<br />

and thus reduce the amount available for disinfection. Small solid particles<br />

suspended in the water also inhibit the action of the UV light since they can<br />

shade or absorb bacteria and also scatter or absorb the light. An advantage of<br />

UV disinfection technology is that small units can be employed to serve small<br />

population bases, whether in the developed or developing world, so the continuous<br />

monitoring activity of chemical systems is avoided. As discussed<br />

later, UV light can also be used to purify water of dissolved organic compounds,<br />

but by a different mechanism.<br />

Disinfection by Chemical Methods:<br />

Ozone and Chlorine Dioxide<br />

To rid drinking water of harmful bacteria and viruses, especially those arising<br />

from human and animal fecal matter, by use of a chemical agent requires an<br />

oxidizing agent more powerful than O2 . In some localities, particularly in<br />

France and other parts of western Europe but also in some North American<br />

cities—Montreal and Los Angeles are examples—ozone is used for this purpose.<br />

Since O3 cannot be stored or shipped because of its very short lifetime,<br />

it must be generated on-site by a relatively expensive process involving electrical<br />

discharge (20,000 V) in dry air. The resulting ozone-laden air is bubbled<br />

through the raw water; about 10 minutes of contact is usually sufficient for<br />

disinfection. Since the lifetime of ozone molecules is short, there is no residual<br />

protection in the purified water to ensure that it will not be subject to<br />

future contamination. Some pollutants in water react with the ozone itself<br />

and others with free radicals such as hydroxyl and hydroperoxy (Chapters<br />

1–5) that are produced when ozone reacts with water.<br />

Unfortunately, the reaction of ozone with bromine in water leads to the<br />

formation of oxygen-containing organic compounds, particularly those con-<br />

taining the carbonyl group, C"O , such as formaldehyde and other low-<br />

molecular-weight aldehydes and various other compounds, some of which are<br />

toxic. In addition, ozone reacts with bromide ion, Br , present in the water<br />

to produce the bromate ion, BrO 3 , a carcinogen in test animals and probably<br />

also in humans. The reaction of ozone with bromide, a natural constituent<br />

of water that is often present at ppm concentrations, occurs in several<br />

steps; the overall reaction is<br />

Br 3 O 3 9: BrO 3 3 O2<br />

The bromate ion produced by ozonation may subsequently react with<br />

organic matter in the water to produce toxic organobromine compounds,<br />

though experiments have shown that the only brominated product under<br />

90


water treatment conditions is dibromoacetonitrile, CHBr 2 CN, produced by<br />

the reaction of bromate ion with acetonitrile. The MCL (maximum contaminant<br />

level) of bromate ion in drinking water is set at 10 ppb (0.010 ppm) by<br />

the U.S. EPA. Substances such as bromate ion that are produced during water<br />

purification are called disinfection by-products, or DBPs. All known chemical<br />

methods of disinfecting water produce DBPs of one type or another.<br />

Similarly, chlorine dioxide gas, ClO 2 , is used in more than 300 North<br />

American and several thousand European communities to disinfect water.<br />

The ClO 2 molecules, themselves free radicals, operate to oxidize organic<br />

molecules by extracting electrons from them:<br />

ClO 2 4 H 5 e 9: Cl 2 H 2 O<br />

The organic cations created in the accompanying oxidation half-reaction<br />

subsequently react with oxygen and eventually become more fully oxidized.<br />

Since chlorine dioxide is not a chlorinating agent—it does not generally<br />

introduce chlorine atoms into the substances with which it reacts—and since<br />

it oxidizes the dissolved organic matter, much smaller amounts of toxic<br />

organic chemical by-products are formed than if molecular chlorine were<br />

used (see below).<br />

As is the case with ozone, ClO 2 cannot be stored since it is explosive in<br />

the high concentrations that its practical use calls for, so it must be generated<br />

on-site. This is accomplished by oxidizing its reduced form, the chlorite ion,<br />

<br />

ClO2 , from the salt sodium chlorite, NaClO2 :<br />

<br />

ClO2 9: ClO2 e <br />

Some of the chlorine dioxide in these processes is converted to chlorate ions,<br />

ClO 3<br />

. The presence of chlorite and chlorate ions as residuals in the final<br />

water has raised health concerns due to their potential toxicity. The U.S.<br />

EPA has set an MCL of 1.0 ppm for chlorite ion, and a MRDL (maximum<br />

residual disinfectant level) of 0.8 ppm for chlorine dioxide, in drinking water.<br />

Disinfection by Chlorination: History<br />

The most common water purification agent used in North America is<br />

hypochlorous acid, HOCl. About half the U.S. population uses surface<br />

water, and one-quarter of the population uses groundwater, that is disinfected<br />

by HOCl. This neutral, covalent compound kills microorganisms, as it readily<br />

passes through their cell membranes. In addition to being effective, disinfection<br />

by chlorination is relatively inexpensive. Incorporating a small excess<br />

of the chemical in the treated water provides it with residual disinfection<br />

power during its subsequent storage and transmission to the consumer. Chlorination<br />

is more common than ozonation in North America because generally<br />

the raw water is less polluted. Chlorination of public water supplies in the<br />

United States, Canada, and Great Britain began in the early years of the<br />

91<br />

Water Disinfection 611


612 Chapter 14 The Pollution and Purification of Water<br />

twentieth century. For the previous 50 years, chlorination had been practiced<br />

on an emergency basis during epidemics caused by water-borne pathogens.<br />

Disinfection by Chlorination: Production<br />

of Hypochlorous Acid<br />

Like ozone, HOCl is not stable in concentrated form and so cannot be stored.<br />

For large-scale installations, e.g., municipal water treatment plants, it is generated<br />

by dissolving molecular chlorine gas, Cl2 , in water. At moderate pH<br />

values, the equilibrium in the reaction of chlorine with water lies far to the<br />

right and is achieved in a few seconds:<br />

Cl2(g) H2O(aq) Δ HOCl(aq) H Cl <br />

Thus a dilute aqueous solution of chlorine in water contains very little aqueous<br />

Cl 2 itself. If the pH of the reaction water were allowed to become too<br />

high, the result would be the ionization of the weak acid HOCl to the<br />

hypochlorite ion, OCl , which is less able to penetrate bacteria on account<br />

of its electrical charge. Once chlorination is complete, the pH of the water is<br />

adjusted upward, if necessary, by the addition of lime, CaO.<br />

In small-scale applications of chlorination, as in swimming pools, the<br />

handling of cylinders of Cl 2 is inconvenient and dangerous. The chlorine gas<br />

can be produced as needed on the spot by the electrolysis of salty water. More<br />

commonly, hypochlorous acid instead is generated from the salt calcium<br />

hypochlorite, Ca(OCl) 2 , or is supplied as an aqueous solution of sodium<br />

hypochlorite, NaOCl. In water, an acid–base reaction occurs to convert most<br />

of the OCl in these substances to HOCl:<br />

OCl H 2O Δ HOCl OH <br />

Close control of the pH in an environment like a swimming pool is necessary<br />

to avoid the shift to the left side in the position of equilibrium for this reaction<br />

that occurs if a very alkaline condition is permitted to prevail. On the<br />

other hand, corrosion of pool construction materials can occur in acidic<br />

water, so the pH is usually maintained above 7 to prevent such deterioration.<br />

Maintenance of an alkaline pH also prevents the conversion of dissolved<br />

ammonia, NH 3, to the chloramines NH 2Cl, NHCl 2, and especially NCl 3,<br />

which is a powerful eye irritant:<br />

NH 3 3 HOCl 9: NCl 3 3 H 2 O<br />

Significant respiratory and eye irritation problems from exposure to chloramines<br />

in the air around indoor swimming pools has been reported when<br />

appropriate ventilation is unavailable.<br />

It is desirable to adjust the equilibrium point in the OCl 9: HOCl<br />

reaction to favor the predominance of the disinfectant molecular species,<br />

HOCl. Since the equilibrium between HOCl and OCl shifts rapidly in<br />

favor of the ion between pH values of 7 and 9, however, the acidity level must<br />

92


e meticulously controlled. Swimming pool acidity can be adjusted by the<br />

addition of acid (in the form of sodium bisulfate, NaHSO 4 , which contains the<br />

acid HSO 4 ) or a base (sodium carbonate, Na2 CO 3 ) or a buffer (sodium<br />

bicarbonate, NaHCO 3, which contains the amphoteric anion HCO 3 ).<br />

Chlorine must be constantly replenished in outdoor pools since UV-B and<br />

the short-wavelength components of UV-A light in sunshine are absorbed<br />

by and decompose the hypochlorite ion, thereby affecting the equilibrium in<br />

the OCl 9: HOCl process toward the ion:<br />

UV<br />

2 ClO 9: 2 Cl O 2<br />

Hypochlorous acid can also be generated by the reaction with water of<br />

the chlorine-containing compound isocyanuric acid, C 3 N 3 O 3 H 3 :<br />

H<br />

O<br />

N N<br />

Either the trichloro derivative, in which each hydrogen is replaced by Cl to<br />

give C 3 N 3 O 3 Cl 3 , or the sodium dichloro derivative, C 3 N 3 O 3 Cl 2 Na, is used.<br />

In either case, the OH group from water combines with the chlorine to<br />

produce HOCl and the hydrogen of H 2 O becomes bonded to the nitrogen,<br />

giving isocyanuric acid, C 3 N 3 O 3 H 3 :<br />

C 3 N 3 O 3 Cl 3 3 H 2 O Δ C 3 N 3 O 3 H 3 3 HOCl<br />

Since this process is an equilibrium, not all the compound is immediately<br />

converted to hypochlorous acid. As HOCl is used up, both by its use as a disinfectant<br />

and through dissociation in sunlight of its ionic form, the equilibrium<br />

shifts to the right and more HOCl is produced. None of the various<br />

forms of isocyanuric acid absorb UV light, so its chlorine is “protected”<br />

against decomposition by sunlight. Since the bulk chlorinated forms of isocyanuric<br />

acid are expensive, it is common to supply hypochlorite from a<br />

cheaper source and to add isocyanuric acid as a stabilizer, temporarily reversing<br />

the above reaction to “store” the chlorine until it is needed.<br />

Disinfection by Chlorine: By-Products and<br />

Their Health Effects<br />

An important drawback to the use of chlorination in disinfecting water is the<br />

concomitant production of chlorinated organic substances, some of which<br />

are toxic, since HOCl is not only an oxidizing agent but also a chlorinating<br />

agent. Examples of these important by-products are the group of halogenated<br />

H<br />

O N O<br />

H<br />

93<br />

Water Disinfection 613


614 Chapter 14 The Pollution and Purification of Water<br />

acetic acids (haloacetic acids), such as CH 2 Cl9COOH, which the U.S. EPA<br />

restricts to 60 ppb as an MCL annual average for drinking water, and haloacetonitriles,<br />

such as CH 2 Cl9CN. Dichloroacetic acid, CHCl 2 9COOH, is<br />

a more potent carcinogen than is chloroform.<br />

If the water to be disinfected contains phenol, C 6 H 5 OH, or a derivative<br />

thereof, chlorine readily substitutes for the hydrogen atoms on the ring to<br />

give rise to chlorinated phenols: These compounds have an offensive odor<br />

and taste and are toxic. Some communities switch from chlorine to chlorine<br />

dioxide when their supply of raw water is temporarily contaminated with<br />

phenols to avoid the formation of chlorinated phenols.<br />

A more general problem with chlorination of water lies in the production<br />

of trihalomethanes, THMs. Their general formula is CHX 3, where the<br />

three X atoms can be chlorine or bromine or a combination of the two. The<br />

THM of principal concern is chloroform, CHCl 3 , which is produced when<br />

hypochlorous acid reacts with organic matter dissolved in the water (see<br />

Box 14-3). Chloroform is a suspected liver carcinogen in humans, and it<br />

may also give rise to negative reproductive and developmental effects. Its<br />

BOX 14-3 The Mechanism of Chloroform Production in Drinking Water<br />

Humic acids, with which HOCl reacts<br />

to form chloroform, are water-soluble,<br />

nonbiodegradable components of decayed<br />

plant matter. Of particular importance are<br />

humic acids that contain 1,3-dihydroxybenzene<br />

rings. The carbon atom (#2) located between<br />

those carrying the 9OH groups is readily<br />

chlorinated by HOCl, as in this elementary<br />

case:<br />

OH O<br />

OH<br />

HOCl<br />

Subsequently the ring cleaves between C-2<br />

and C-3 to yield a chain:<br />

R9C9CHCl 2<br />

O<br />

1<br />

2<br />

3<br />

Cl<br />

Cl<br />

OH<br />

In the presence of the HOCl, the terminal carbon<br />

becomes trichlorinated, and the 9CCl 3<br />

group is readily displaced by the OH in water<br />

to yield chloroform:<br />

R9C9CHCl 2<br />

O<br />

HOCl<br />

R9C9CCl 3<br />

Analogous sequences of reactions produce<br />

bromoform, CHBr 3, and mixed chlorine–<br />

bromine trihalomethanes from the action on<br />

humic materials of hypobromous acid, HOBr,<br />

which is formed when bromide ion in water<br />

displaces chlorine from HOCl:<br />

94<br />

OH <br />

H <br />

O<br />

R9C9OH CHCl 3<br />

O<br />

HOCl Br EF HOBr Cl


presence, even at very low levels of approximately 30 ppb, raises the specter<br />

that chlorinated drinking water may pose a health hazard, though one that<br />

pales by comparison with the benefits that it confers in the elimination of<br />

fatal waterborne diseases. The annual average limit of total THMs in drinking<br />

water in the United States and the European Union (EU) has been<br />

reduced to 80 ppb. The previous limit of 100 ppb is still used in Canada. In<br />

fact, these 80–100-ppb limits are set not only to regulate the THM chemicals<br />

themselves but also as an indicator that the production of other chlorinated<br />

organic DBPs (see below) is not excessive.<br />

The U.S. EPA has set a maximum contaminant level goal, MCLG, of<br />

70 ppb for THMs in drinking water. An MCLG is the maximum level at which<br />

the contaminant is believed to be safe, allowing for adequate margins of safety, but<br />

unlike the MCL (Chapter 10), it is not an enforceable standard. A nonzero<br />

goal is considered by some scientists and policymakers to be appropriate to<br />

substances, like chloroform, that are believed to operate indirectly as carcinogens.<br />

They do not damage DNA directly but cause tissue damage that leads to<br />

rapid cell proliferation, which in turn increases the likelihood that cancer will<br />

form in the damaged tissue. A threshold below which no effects are likely to<br />

be observed is expected for carcinogens that operate in this manner.<br />

The level of trihalomethanes formed in water depends sharply on the<br />

organic content of the raw water since the THMs are formed from the reaction<br />

of organics with HOCl (see Box 14-3). THM levels can reach 250 ppb<br />

in areas of Scotland and Northern Ireland that have peat moorlands. Water<br />

exposed to bogs in Newfoundland, Canada, has generated THM levels in<br />

excess of 400 ppb. As of the early 1990s, about 1% of the larger U.S. drinking<br />

water utilities that used surface waters, but none that used groundwater, had<br />

average THM levels exceeding 100 ppb. The THM content of chlorinated<br />

water could be decreased by using activated carbon either to remove dissolved<br />

organic compounds before the water is chlorinated or to remove<br />

THMs and other chlorinated organics after the process, although THMs are<br />

not very efficiently adsorbed by the carbon and it is an expensive process.<br />

An analysis of epidemiological studies relating the chlorination of water<br />

to cancer rates in various communities in the United States led to the conclusion<br />

that the risk of bladder cancer in humans increased by 21%, and that<br />

of rectal cancer by 38%, for Americans who drank chlorinated surface water<br />

in the past. A similar study in Ontario found even higher bladder cancer risk<br />

factors for people who drank water for 35 years or more that had THM levels<br />

greater than 50 ppb, and for colon cancer when the concentration exceeded<br />

75 ppb, but found no correlations of THMs with rectal cancer rates. A recent<br />

study found no increased risk for pancreatic cancer from lifetime exposure to<br />

the by-products of chlorinated water.<br />

Given that slightly more than half the population of the United States<br />

drinks surface water, one effect of chlorination is to have increased bladder<br />

cancer incidence by about 4200 cases per year and rectal cancer incidence by<br />

95<br />

Water Disinfection 615


616 Chapter 14 The Pollution and Purification of Water<br />

about 6500 cases annually. Because of these risks, some communities are considering<br />

a switch, or have already switched, to water disinfection by ozone or<br />

chlorine dioxide, since these agents produce little or no chloroform. The<br />

extent of chlorination has already been reduced in most American communities<br />

relative to the levels that led to these statistics.<br />

Several other mutagenic chlorinated organic DBPs formed during chlorination<br />

have been detected in water, in addition to chloroform. It is not clear<br />

whether the main carcinogens in the chlorinated drinking water are the<br />

THMs themselves or some nonvolatile, higher-molecular-weight mutagenic<br />

by-product present at still lower concentrations but whose concentration<br />

would presumably be proportional to THM. The same risks do not usually<br />

apply to chlorinated well water, since its organochlorine content is much less<br />

(only 0.8 ppb on average, versus 51 ppb for surface water) because it contains<br />

much smaller amounts of organic matter that could become chlorinated.<br />

Exposure to chloroform by dermal contact and inhalation of the gases deabsorbed<br />

from the hot water during showers, baths, and hand-washing of dishes<br />

contribute more to one’s intake of THMs, as measured by the blood levels of<br />

these compounds, than does drinking the water itself. Swimming in pools in<br />

which the water is chlorinated for disinfection also contributes significantly<br />

to dermal exposure. Recently it has been found that use of the disinfectant<br />

triclosan, a phenol-based compound present in some hand soaps, in chlorinated<br />

water can produce additional chloroform and chlorinated phenols, to<br />

which the user is then exposed.<br />

Recently, public health officials have expressed concern about the possible<br />

link between THMs and adverse human reproductive outcomes, including<br />

first-term miscarriages, stillbirths, impaired fetal growth, and certain birth<br />

defects. Even though the existing research in this area is not yet definitive,<br />

some officials suggest that women drink bottled water rather than chlorinated<br />

tap water during their first three months of pregnancy.<br />

Disinfection by Chlorine: Advantages over Other Methods<br />

Notwithstanding the preceding discussion of chlorination by-products, it is<br />

important to point out that the disinfection of water by chlorine is extremely<br />

important in protecting public health and saves many more lives—by a very<br />

wide margin—than are affected negatively. For example, both typhoid and<br />

cholera were widespread in both Europe and North America a century ago<br />

but have been almost completely eradicated in the developed world, thanks<br />

to chlorination and the other disinfection methods for drinking water and to<br />

improved sanitation in general. The same is not true in many developing<br />

countries; e.g., there were more than half a million cases of cholera in Peru in<br />

the early 1990s. Overall, about 20 million people, most of them infants, die<br />

from waterborne diseases annually worldwide in underdeveloped countries,<br />

where water purification is often erratic or even nonexistent. Under no<br />

96


circumstances should effective disinfection of water be abandoned because of<br />

concern for the by-products of chlorination!<br />

An advantage chlorination has over disinfection by chlorine dioxide or<br />

ozone or by UV is that some chlorine remains dissolved in water once it has<br />

left the purification plant, so that the water is protected from subsequent bacterial<br />

contamination before it is consumed. Indeed, some chlorine is usually<br />

added to water purified by the other methods to provide this protection.<br />

There is very little danger of significant chloroform production in the purified<br />

water since its organic content has been virtually eliminated before the chlorine<br />

is introduced. If the chlorine level in water purified by chlorination is too<br />

high, it can be lowered by the addition of sulfur dioxide.<br />

The residual chlorine in water often exists in the form of the chloramines<br />

NH 2 Cl, NHCl 2 , and NCl 3 , which are produced from reaction with dissolved<br />

ammonia gas, NH 3 . Although not as fast as HOCl in disinfecting water, the<br />

mono- and dichloroamines especially are good disinfectants. The mixture of<br />

chloramines, called combined chlorine, is longer-lived in water than is<br />

hypochlorous acid and thus provides longer residual protection. Indeed,<br />

ammonia is often added to purified drinking water in order to convert the<br />

residual chlorine to the combined form (Figure 14-1). Chloramines are sometimes<br />

used, rather than chlorine or ozone or chlorine dioxide, as the main disinfectant<br />

in the purification of drinking water. They have the advantage over<br />

chlorine of producing little (though not zero) amounts of THMs and<br />

haloacetic acids. The U.S. EPA has set MRDLs of 4.0 ppm for both chlorine<br />

and chloramine in drinking water.<br />

Bromine rather than chlorine is sometimes used as the disinfectant in<br />

swimming pools. The main disinfecting agent in bromination is hypobromous<br />

acid, HOBr, in analogy with the role of hypochlorous acid in chlorination.<br />

HOBr reacts more rapidly with dissolved ammonia than does HOCl,<br />

producing mainly NH 2 Br, which is also a good disinfectant.<br />

To disinfect water for drinking purposes, hikers either boil raw water or<br />

treat it chemically with either chlorine, in the form of bleach, which provides<br />

HOCl, or iodine, as elemental I 2 or hypoiodous acid, HIO. Concerns have<br />

been expressed about chronic health problems such as thyroid dysfunction<br />

associated with long-term usage of iodine, however. Treating the water with<br />

elemental iodine tends to make it unpalatable as well.<br />

PROBLEM 14-3<br />

Assuming that the nitrogen atom in monochloramine, NH 2 Cl, has an oxidation<br />

number of 3, calculate that of the chlorine atom. Using the principle<br />

that unlike charges attract, predict whether it will be the hydrogen ion or the<br />

hydroxide ion from dissociated water molecules that will extract the Cl from<br />

NH 2 Cl; from your result, predict the products of the decomposition reaction<br />

of chloramine in water.<br />

97<br />

Water Disinfection 617


618 Chapter 14 The Pollution and Purification of Water<br />

Surface<br />

Water table<br />

FIGURE 14-3 Groundwater<br />

location in relation<br />

to regions in the soil.<br />

Precipitation<br />

Soil moisture<br />

Aeration zone<br />

(unsaturated)<br />

Groundwater in<br />

saturated zone<br />

(aquifer)<br />

Clay or impervious rock<br />

A drinking-water quality issue of current concern involves the pathogenic<br />

protozoa called Cryptosporidium, which was responsible for the death of<br />

100 people and for about 400,000 cases of watery diarrhea in Milwaukee in<br />

1993. Less serious Cryptosporidium outbreaks occurred in Oxford, England, in<br />

1989, and in Saskatchewan, Canada, in 2001. This deadly parasite is resistant<br />

to standard methods of disinfection, such as chlorination at normal levels,<br />

and is so small (about 5-mm diameter) that it easily passes through the standard<br />

filters used to separate sediments. Several possible solutions have been<br />

advanced, including ozonation or the use of ultrafiltration or UV irradiation<br />

or the application of monochloramine following chlorination. A longerthan-usual<br />

exposure of water containing Cryptosporidium is necessary with<br />

ozonation, since the activation energy for inactivation of protozoa by ozone<br />

is about twice as large as that for bacteria (80 vs. about 40 kJ/mol).<br />

Another protozoa, Giardia lamblia, also causes many instances of waterborne<br />

disease. Like Cryptosporidium, it is also somewhat resistant to chlorination,<br />

but since it is larger (about 10-mm diameter), it is more easily removed<br />

by filtration through sand.<br />

Groundwater: Its Supply, Chemical<br />

Contamination, and Remediation<br />

The Nature and Supply of Groundwater<br />

The great majority of the available fresh water on Earth lies underground,<br />

half of it at depths exceeding a kilometer. As one digs into the ground below<br />

the initial belt of soil moisture, the aeration or unsaturated zone, where the<br />

particles of soil are covered with a film of water but in<br />

which air is present between the particles, is next<br />

encountered. At lower depths is the saturated zone, in<br />

which water has displaced all the air from these pore<br />

spaces. Groundwater is the name given to the fresh<br />

water in the saturated zone (see Figure 14-3); it makes<br />

up 0.6% of the world’s total water supply. The ultimate<br />

source of groundwater is precipitation that falls onto<br />

the surface; a small fraction of it eventually filters<br />

down to the saturated zone. Underground water ranges<br />

in “age” from a few years to millions of years. For<br />

example, in zones that are now arid, much of the<br />

groundwater currently being accessed has been present<br />

since the wetter conditions of the last ice age and will<br />

not be quickly replaced.<br />

The top of the groundwater (saturated) region is<br />

called the water table. In some places it occurs right at<br />

the surface of the soil, a phenomenon that gives rise to swamps. Where the<br />

water table lies above the soil, we encounter lakes and streams.<br />

98


Groundwater: Its Supply, Chemical Contamination, and Remediation 619<br />

If groundwater is contained in soil that is composed of porous rocks such<br />

as sandstone, or in highly fractured rock such as gravel or sand, and if the<br />

water is bounded at its lower depths by a layer of clay or impervious rocks,<br />

then it constitutes a permanent reservoir—a sort of underground lake—<br />

called an aquifer. Some aquifers lie below several layers of impermeable rock<br />

or soil; these are called confined or artesian aquifers.<br />

Groundwater in aquifers can be extracted by wells, and it is the main supply<br />

of drinking water for almost half the population of North America and<br />

over 1.5 billion people worldwide. In the United States in 1990, groundwater<br />

supplied 39% of the water used for public supplies and 96% of that withdrawn<br />

for individual domestic systems, the latter being very common in rural homes.<br />

In Europe the proportion of public drinking water extracted from aquifers<br />

ranges from nearly 100% for Denmark, Austria, and Italy, to about two-thirds<br />

in Germany, Switzerland, and the Netherlands, to less than one-third in<br />

Great Britain and Spain.<br />

In the United States, the majority of groundwater usage is for irrigation<br />

purposes, almost all of it in the western states. The massive extraction of<br />

water from American aquifers has given rise to fears about future supplies of<br />

fresh water (and about the sinking of land above the aquifers), since such<br />

aquifers are replenished only very slowly. In the High Plains of the central<br />

United States, more than half the groundwater in storage has been depleted<br />

in some areas. In northern China, the depletion of shallow aquifers is forcing<br />

the sinking of wells more than 1 km deep in order to reach a new supply<br />

of groundwater. Indeed, groundwater depletion—along with the buildup<br />

of salts in the soil—is now the dominant threat to irrigated agriculture. In<br />

addition, the contamination of groundwater by chemicals is becoming a<br />

serious concern in many areas. A side effect of the rise in sea levels that will<br />

accompany global warming is the intrusion of salt water into aquifers near<br />

coasts.<br />

The Contamination of Groundwater<br />

Groundwater has been traditionally considered to be a pure form of water.<br />

Because of its filtration through soil and its long residence time underground,<br />

it contains much less natural organic matter and many fewer disease-causing<br />

microorganisms than water from lakes or rivers, although the latter point may<br />

be a misconception, according to recent evidence. Some groundwater is naturally<br />

too salty or too acidic for either drinking or irrigation purposes and<br />

may contain too much sodium, sulfide, or iron for many uses.<br />

Humans have been concerned about the pollution of surface water in<br />

rivers and lakes for a long time. Indeed, a recent survey indicated that stream<br />

water in both agricultural and urban areas of the United States contains pesticide<br />

concentrations that exceed human-health benchmark standards. In<br />

contrast, the contamination by chemicals of groundwater was not recognized<br />

as a serious environmental problem until the 1980s, notwithstanding the fact<br />

99


620 Chapter 14 The Pollution and Purification of Water<br />

that it had been occurring for half a century. To a large extent, groundwater<br />

contamination was neglected because it was not immediately visible—it was<br />

“out of sight, out of mind”—even though groundwater is a major source of<br />

drinking water. We were ignorant of the long-range consequences of our<br />

waste disposal practices. Ironically, surface water can be cleaned up relatively<br />

easily and quickly, whereas groundwater pollution is a much harder, much<br />

more expensive, long-range problem to solve.<br />

Because we are now aware of the consequences—including high remediation<br />

costs—of the uncontrolled disposal of organic chemical wastes, most<br />

large corporations in developed countries have become much more responsible<br />

in their disposal of chemicals. Unfortunately, the collective discharges<br />

from smaller sources, including many municipalities, small industries, and<br />

farms, have not yet been controlled in like manner. Similarly, the huge number<br />

of septic tanks that exist are collectively a major source of nitrate, bacteria,<br />

viruses, detergents, and household cleaners to groundwater.<br />

Nitrate Contamination of Groundwater<br />

The inorganic contaminant of greatest concern in groundwater is the nitrate<br />

<br />

ion, NO3 , which commonly occurs in both rural and suburban aquifers.<br />

Although uncontaminated groundwater generally has nitrate nitrogen levels<br />

of 4–9 ppm, about 9% of shallow aquifers—from which water is often<br />

extracted via privately owned wells—in the United States now have nitrate<br />

levels that exceed the 10-ppm nitrogen MCL value. Indeed, elevated levels<br />

of about 100 ppm can result from agricultural activity. The location of areas<br />

in the United States that have a high risk of nitrate contamination of<br />

groundwater is shown in Figure 14-4. Exceeding the 10-ppm MCL limit is<br />

much rarer (1%) for public U.S. groundwater supplies, partially because they<br />

are drawn from deeper aquifers; these are generally less contaminated because<br />

of their depth, because their location is remote from large sources of contamination,<br />

and because natural remediation via denitrification of nitrate in the<br />

low-oxygen conditions can occur.<br />

The expenditure of public money on nitrate-level reductions in drinking<br />

water has become a controversial subject. In Great Britain, in particular, hundreds<br />

of millions of dollars have been spent on achieving the 50-ppm maximum<br />

level of nitrate ion set by the European Union. Because nitrate removal<br />

from well water is very expensive, water contaminated with high levels of the<br />

ion are not normally used for human consumption, at least in public supplies.<br />

PROBLEM 14-4<br />

Convert the EU nitrate standard of 50 ppm to its nitrogen content alone.<br />

Is the EU standard more or less stringent than the U.S. regulatory limit of<br />

10 ppm nitrogen as nitrate per liter?<br />

100


Increasing risk of<br />

groundwater<br />

contamination<br />

Groundwater: Its Supply, Chemical Contamination, and Remediation <strong>621</strong><br />

Nitrogen<br />

input<br />

High<br />

High<br />

Low<br />

Low<br />

Aquifer<br />

vulnerability<br />

High<br />

Low<br />

High<br />

Low<br />

Nitrate in groundwater originates mainly from four sources:<br />

• application of nitrogen fertilizers, both inorganic and animal manure, to<br />

cropland,<br />

• cultivation of the soil,<br />

• human sewage deposited in septic systems, and<br />

• atmospheric deposition.<br />

Concern has been expressed about the increasing levels of nitrate ion in<br />

drinking water, particularly in well water in rural locations; the main source<br />

of this NO 3 is runoff from agricultural lands into rivers and streams. Almost<br />

12 million tons of nitrogen is applied annually as fertilizer for agriculture in<br />

the United States, and manure production contributes almost 7 million tons<br />

more. Initially, oxidized animal wastes (manure), unabsorbed ammonium<br />

nitrate, NH 4 NO 3 , and other nitrogen fertilizers were thought to be the culprits<br />

in nitrate contamination of groundwater, since reduced nitrogen unused<br />

by plants is converted naturally to nitrate, which is highly soluble in water<br />

101<br />

FIGURE 14-4 The risk of<br />

nitrate contamination of the<br />

groundwaters in the United<br />

States. [Source: B. T. Nolan<br />

et al., “Risk of Nitrate in<br />

Groundwaters of the<br />

United States—A National<br />

Perspective,” Environmental<br />

Science and Technology 31<br />

(1997): 2229.]


622 Chapter 14 The Pollution and Purification of Water<br />

and can easily leach down into groundwater. It now appears that intensive<br />

cultivation of land, even without the application of fertilizer or manure, also<br />

facilitates the oxidation of reduced nitrogen to nitrate in decomposed organic<br />

matter in the soil by providing aeration and moisture. The original, reduced<br />

forms of nitrogen become oxidized in the soil to nitrate, which, being mobile,<br />

then migrates down to the groundwater, where it dissolves in water and is<br />

diluted. Denitrification of nitrate to nitrogen gas (see Chapter 6) and uptake<br />

of nitrate by plants can occur in forested areas that separate agricultural farms<br />

from streams, thereby lowering the risk of contamination in areas with significant<br />

woodland. However, rural areas with high nitrogen input, well-drained<br />

soil, and little woodland are at particular risk for nitrate contamination of<br />

groundwater.<br />

The atmospheric deposition of nitrate results from its production in the<br />

atmosphere when NO X emissions from vehicles and power plants, and its<br />

natural sources in thunderstorms, are oxidized in air to nitric acid and then<br />

neutralized to ammonium nitrate (see Chapters 3 and 5).<br />

In urban areas, the use of nitrogen fertilizers on domestic lawns and golf<br />

courses, parks, etc. contributes nitrate to groundwater. Septic tanks and<br />

cesspools also are significant contributors where they exist.<br />

Excess nitrate ion in wastewater flowing into seawater, e.g., the Baltic<br />

Sea, has resulted in algal blooms that pollute the water after they die. Nitrate<br />

ion normally does not cause this effect in bodies of fresh water, where phosphorus<br />

rather than nitrogen is usually the limiting nutrient; increasing the<br />

nitrate concentration there without an increase in phosphate levels does not<br />

lead to an increased amount of plant growth. There are, however, instances<br />

where nitrogen rather than phosphorus temporarily becomes the limiting<br />

nutrient even in fresh waters.<br />

PROBLEM 14-5<br />

The nitrate concentration in an aquifer is 20 ppm, and its volume is 10 million<br />

liters. What mass of ammonia upon oxidation would have produced this<br />

mass of nitrate?<br />

Health Hazards of Nitrates in Drinking Water<br />

Excess nitrate ion in drinking water is a potential health hazard since it can<br />

result in methemoglobinemia in newborn infants as well as in adults with a<br />

specific enzyme deficiency. The pathological process, in brief, runs as follows.<br />

Bacteria, e.g., in unsterilized milk-feeding bottles or in the baby’s stomach,<br />

reduce some of the nitrate to nitrite ion, NO 2 :<br />

NO 3 2 H 2 e 9: NO2 H2O<br />

The nitrite combines with and oxidizes the iron ions in the hemoglobin in<br />

blood from Fe 2 to Fe 3 and thereby prevents the proper absorption and<br />

102


Groundwater: Its Supply, Chemical Contamination, and Remediation 623<br />

transfer of oxygen to cells. The baby turns blue and suffers respiratory failure.<br />

(In almost all adults, the oxidized hemoglobin is readily reduced back to its<br />

oxygen-carrying form, and the nitrite is readily oxidized back to nitrate; also,<br />

nitrate is mainly absorbed in the digestive tract of adults before reduction to<br />

nitrite can occur.) Methemoglobinemia, or blue-baby syndrome, is now relatively<br />

rare in industrialized countries. It was a serious problem in Hungary up<br />

until the late 1980s and in Romania.<br />

The U.S. EPA MCL of 10 ppm of nitrate nitrogen was set in order to<br />

avoid blue-baby syndrome. Since the syndrome is now almost nonexistent in<br />

the United States (only two cases since the mid-1960s), some policy analysts<br />

think this value is too stringent.<br />

Recently, an increase in the risk of acquiring non-Hodgkin’s lymphoma<br />

has been found for persons in some communities in Nebraska who consume<br />

drinking water having the highest levels (long-term average of 4 ppm or more<br />

of nitrogen as nitrate) of nitrate. As discussed in the next section, excess<br />

nitrate ion in drinking water is also of concern because of its potential link<br />

with stomach cancer. Recent epidemiological investigations have, however,<br />

failed to establish any positive, statistically significant relationship between<br />

nitrate levels in drinking water and the incidence of stomach cancer. A study<br />

reported in 2001 found that older women in Iowa who drank water from<br />

municipal supplies having elevated nitrate levels ( 2.46 ppm) were almost<br />

three times as likely to be diagnosed with bladder cancer than those least<br />

exposed ( 0.36 ppm in their drinking water). However, a recent large-scale<br />

study from the Netherlands failed to find an association between nitrate<br />

exposure and the risk of bladder cancer. A review of the current literature<br />

concluded that there is also no association between nitrate exposure from<br />

drinking water and adverse reproductive effects.<br />

Nitrosamines in Food and Water<br />

Some scientists have warned that excess nitrate ion in drinking water and<br />

foods could lead to an increase in the incidence of stomach cancer in<br />

humans, since some of it is converted in the stomach to nitrite ion. The<br />

nitrites could subsequently react with amines to produce N-nitrosamines,<br />

compounds that are known to be carcinogenic in animals. N-nitrosamines<br />

are amines in which two organic groups and an 9N"O unit are bonded to<br />

the central nitrogen:<br />

R<br />

R<br />

N9N"O<br />

H 3 C<br />

H 3 C<br />

N9N"O<br />

N-nitrosamines NDMA<br />

103


624 Chapter 14 The Pollution and Purification of Water<br />

Of concern not only with respect to its production in the stomach and its<br />

occurrence in foods and beverages (e.g., cheeses, fried bacon, smoked and/or<br />

cured meat and fish, and beer), but also as an environmental pollutant in<br />

drinking water, is the compound in which R in the above structure is the<br />

methyl group CH 3 ; it is called N-nitrosodimethylamine, or NDMA for short.<br />

This organic liquid is somewhat soluble in water (about 4 g/L) and somewhat<br />

soluble in organic liquids. It is a probable human carcinogen, and a potent<br />

one if extrapolation from animal studies is a reliable guide. It can transfer a<br />

methyl group to a nitrogen or oxygen of a DNA base, thereby altering the<br />

instructional code for protein synthesis in the cell.<br />

In the early 1980s, it was found that NDMA was present in beer to the<br />

extent of about 3000 ppt. Since that time, commercial brewers have modified<br />

the drying of malt so that the current levels of NDMA in American and<br />

Canadian beers are now only about 70 ppt.<br />

Large quantities of nitrate are used to “cure” pork products such as bacon<br />

and hot dogs. In these foods, some of the nitrate ion is biochemically reduced<br />

to nitrite ion, which prevents the growth of the organism responsible for<br />

botulism. Nitrite ion also gives these meats their characteristic taste and color<br />

by combining with hemoproteins in blood. Nitrosamines are produced from<br />

excess nitrite during frying (e.g., of bacon) and in the stomach, as discussed.<br />

Government agencies have instituted programs to decrease the residual nitrite<br />

levels in cured meats. Some manufacturers of these foods now add vitamin C<br />

or E to the meat in order to block the formation of nitrosamines. Based upon<br />

average levels of NDMA in various foods and the average daily intake for each<br />

of them, most of us now ingest more NDMA from consumption of cheese<br />

(which is often treated with nitrates) than from any other source.<br />

Even though the commercial production of NDMA has been phased out,<br />

it can be formed as a by-product due to the use of amines in industrial<br />

processes such as rubber tire manufacturing, leather tanning, and pesticide<br />

production.<br />

The levels of NDMA in drinking water drawn from groundwater is of<br />

concern in some localities that have industrial sources of the compound. For<br />

example, following the discovery that the water supply of one town had been<br />

contaminated by up to 100 ppt NDMA from a tire factory, Ontario, Canada,<br />

adopted a guideline maximum of 9 ppt of NDMA in drinking water, which<br />

corresponds to a lifetime cancer risk of 1 in 100,000. By contrast, the guideline<br />

for water in the United States is set at 0.68 ppt, which corresponds to a<br />

cancer risk of 1 in a million, but which actually lies considerably below the<br />

detection limit (about 5 ppt) for the compound.<br />

PROBLEM 14-6<br />

Write balanced redox half-reactions (assuming acidic conditions) for the<br />

<br />

conversion of NH4 to NO3 , and of NO2 to N2.<br />

104


Perchlorates<br />

Groundwater: Its Supply, Chemical Contamination, and Remediation 625<br />

Perchlorate ion, ClO 4 , is analogous to nitrate ion in that both oxyanions<br />

involve nonmetals in their highest common oxidation numbers (7 inthe<br />

case of perchlorate; see Additional Problem 13 to generate a summary of<br />

chlorine with its many oxidation numbers). For that reason, both ions are<br />

oxidizing agents, and both have been used in explosives and propellants. Both<br />

have both natural and anthropogenic sources. Perchlorate is a newly discovered<br />

(late 1990s) pollutant in the drinking-water supply of about 15 million<br />

Americans. Large quantities of ammonium perchlorate, NH 4ClO 4, are manufactured<br />

for use as oxidizing agents in solid rocket propellants, fireworks,<br />

batteries, and automobile air bags. Because rocket fuel has a limited shelf life,<br />

it must be replaced regularly. Large amounts of perchlorates were washed out<br />

of missiles and rocket boosters onto the ground or into holding lagoons in the<br />

second half of the twentieth century.<br />

Perchlorate contamination has been established in 23 U.S. states,<br />

including much of the Colorado River and aquifers in the deserts of the<br />

Southwest. Concentrations of perchlorate in drinking water in the U.S.<br />

Southwest range from 5 to 20 ppb. The map (Figure 14-5) of perchlorate<br />

releases in the United States indicates that most occur in the south-central<br />

and western states, especially California. Perchlorate has also been found in<br />

garden fertilizers at concentrations approaching 1%. It occurs naturally in<br />

some Chilean deposits of nitrate which are exported to the United States and<br />

elsewhere as fertilizers. Perchlorate also exists naturally in some minerals<br />

Confirmed sites<br />

Unconfirmed sites<br />

Urbanized areas<br />

Major rivers<br />

Affected states<br />

105<br />

FIGURE 14-5 Regions<br />

of perchlorate use and<br />

contamination in the<br />

United States. [Source: B. E.<br />

Logan, “Assessing the Outlook<br />

for Perchlorate Remediation,”<br />

Environmental Science and<br />

Technology (1 December<br />

2001): 484A.]


626 Chapter 14 The Pollution and Purification of Water<br />

found in the U.S. Southwest. A recent analysis indicates that both its use as an<br />

oxidizer, in fireworks, rockets, etc., and its presence in fertilizer make important<br />

contributions to its contamination of foodstuffs in the United States.<br />

At high doses, perchlorate affects human health by reducing hormone<br />

production in the thyroid, where it competes with iodide ion. Its hazard<br />

at low concentrations, if any, is not known, making the development of a<br />

drinking-water standard for the ion a difficult problem. No federal MCL for<br />

the ion has yet been set, but several states have set their own limits, ranging<br />

from 1 to 18 ppb; e.g., California’s limit is 6 ppb. In 2002, a draft report by the<br />

U.S. EPA proposed a drinking-water standard of 1 ppb as safe for human<br />

health, but this value has been criticized as too low by the Department of<br />

Defense and companies that make or use perchlorates. Research reported in<br />

2002 on volunteers indicated that the no-effect level (NOEL; Chapter 10) for<br />

the inhibition of iodine uptake corresponds to a drinking-water concentration<br />

of at least 180 ppb. The EPA based its 1-ppb recommendation on studies<br />

indicating that mothers who drink perchlorate-contaminated water could<br />

give birth to children whose IQs would be affected negatively because correct<br />

maternal thyroid hormone levels are vital to fetal brain development.<br />

Like nitrate, perchlorate is a difficult ion to remove from water supplies<br />

since it is a highly water-soluble anion that is very inert and does not adsorb<br />

readily either to mineral surfaces or to activated carbon. Barrier methods,<br />

using elemental iron, etc. are not successful because the anion is so unreactive.<br />

The primary technologies currently in use to remediate perchlorate in<br />

water are ion exchange and biological treatment. Some ion exchange resins<br />

successfully remove perchlorate, though it tends to remain in solution until<br />

all other anions have been absorbed. Ion exchange is used especially when<br />

perchlorate concentrations are low to begin with.<br />

Certain bacteria found naturally in many soils, sediments, and natural<br />

waters biodegrade perchlorate by reduction to chloride ion:<br />

<br />

ClO4 9: 9: Cl 2 O2<br />

Perchlorate can be biodegraded in bioreactors, large vats that are engineered<br />

to maintain a high concentration of the appropriate bacteria in contact with<br />

the water.<br />

Groundwater Contamination by Organic Chemicals<br />

The contamination of groundwater by organic chemicals is a major concern.<br />

Many organic substances decay rapidly or are immobilized in the soil, so the<br />

number of compounds that are sufficiently persistent and mobile to travel to<br />

the water table and to contaminate groundwater there is relatively small.<br />

The compounds that are most often detected in groundwater-based U.S.<br />

community public water supplies, including those near hazardous waste sites,<br />

are summarized in Table 14-1. Municipal landfills as well as industrial waste<br />

106


TABLE 14-1<br />

Groundwater: Its Supply, Chemical Contamination, and Remediation 627<br />

Organic Compounds Commonly Found in U.S. Groundwater-Based<br />

Community Water Supplies and Their Properties<br />

Chemical Density (g/mL) Water Solubility (g/L)<br />

Present at 25–50% of sites:<br />

Chloroform (trichloromethane) 1.48 8.2<br />

Bromodichloromethane 1.98 4.4<br />

Dibromochloromethane 2.45 2.7<br />

Bromoform (tribromomethane)<br />

Present at a smaller fraction of sites:<br />

2.89 3.0<br />

Trichloroethene 1.46 1.1<br />

Tetrachloroethene (perchloroethene) 1.62 0.15<br />

1,1,1-Trichloroethane 1.34 1.5<br />

1,2-Dichloroethenes 1.26, 1.28 3.5, 6.3<br />

1,1-Dichloroethane 1.18 5.5<br />

Carbon tetrachloride 1.46 0.76<br />

Dichloroiodomethane 1.58<br />

Xylenes 0.86–0.88 0.18 (o)<br />

1,2-Dichloropropane 1.16 2.8<br />

Benzene 0.88 1.8<br />

Toluene<br />

Also commonly present at wells close to<br />

hazardous waste sites:<br />

0.87 0.54<br />

Methylene chloride 1.33 20<br />

Ethylbenzene 0.87 0.15<br />

Acetone 0.79 sol<br />

1,1-Dichloroethene 1.22 2.3<br />

1,2-Dichloroethane 1.24 8.5<br />

Vinyl chloride (chloroethene) gas 8.8<br />

Methyl ethyl ketone 0.80 268<br />

Chlorobenzene 1.11 0.47<br />

1,1,2-Trichloroethane 1.44 4.5<br />

Chloroethane 0.90 5.7<br />

Fluorotrichloromethane gas 1.1<br />

1,1,2,2-Tetrachloroethane 1.60 2.7<br />

Methyl isobutyl ketone 0.80 19<br />

Source: Based on U.S. EPA surveys of about 2% of U.S. water supplies.<br />

disposal sites are often the source of the contaminants. Liquid that contains<br />

dissolved matter that drains from a terrestrial source, such as a landfill, is called<br />

a leachate. In rural areas, the contamination of shallow aquifers by organic<br />

pesticides, such as atrazine (Chapter 10) leached from the surface, has become<br />

107


628 Chapter 14 The Pollution and Purification of Water<br />

a concern. The insecticide dieldrin (Chapter 10), which has been banned since<br />

1992, is the pesticide found most often to exceed human-health guideline<br />

levels in U.S. groundwater. Ironically, shallow groundwater aquifers used to<br />

supply drinking water are often more polluted by pesticides at greater than<br />

acceptable levels than are those in agricultural areas in the United States.<br />

The typical organic contaminants in most major groundwater supplies are:<br />

• Chlorinated solvents, especially trichloroethene (TCE, “tric,” also called<br />

trichloroethylene), C2HCl3, and perchloroethene (PCE, “perc,” also called<br />

perchloroethylene or tetrachloroethene), C2Cl4 . These molecules contain a<br />

C"C bond, with three or four of the four hydrogen atoms of ethene (ethylene)<br />

replaced by chlorine:<br />

H<br />

Cl<br />

C"C<br />

Cl<br />

Cl<br />

C"C<br />

By a large margin, chlorinated solvents are the most prevalent organic pollutants<br />

in groundwater.<br />

• Hydrocarbons from the BTX component of gasoline and other petroleum<br />

products: benzene, C6H6 , and its methylated derivatives toluene,<br />

C6H5(CH3), and the three isomers of xylene, C6H4(CH3) 2. (See Chapter 7<br />

for structures.)<br />

• MTBE (methyl tertiary-butyl ether) from gasoline (see Chapter 8)<br />

The chemicals in the groups mentioned above occur commonly in<br />

groundwater at sites where manufacturing and/or waste disposal occurred,<br />

especially from 1940 to 1980. In that period, little attention was paid to the<br />

ultimate fate and residence following the in-ground injection of these chemicals.<br />

The sources of these organic substances also include leaking chemical<br />

waste dumps, leaking underground gasoline storage tanks, leaking municipal<br />

landfills, and accidental spills of chemicals on land.<br />

Trichloroethene is an industrial solvent, used to dissolve grease on metal, as<br />

is perchloroethene. The U.S. MCL for TCE in drinking water is 5 ppb, and the<br />

same limit is now used in Canada as well. A 2006 report by the U.S. National<br />

Academy of Sciences concluded that TCE is a possible cause of kidney cancer,<br />

can impair neurological function, and can cause reproductive and developmental<br />

damage. A link between TCE exposure and an abnormally low sperm count<br />

in males has been established. The International Agency for Research on<br />

Cancer has classified TCE as “probably carcinogenic to humans.”<br />

PCE is used not only in metal degreasing but also finds wide application<br />

as the solvent in dry-cleaning operations, so it is released from a large number<br />

of small sources. A group of women in Cape Cod, Massachusetts, who were<br />

inadvertently exposed over several decades to high levels of PCE in their<br />

Cl<br />

Cl<br />

TCE PCE<br />

108<br />

Cl<br />

Cl


Groundwater: Its Supply, Chemical Contamination, and Remediation 629<br />

drinking water were found to have small to moderate increases in their risk of<br />

contracting breast cancer.<br />

Gasoline enters the soil via surface spills, leakage from underground storage<br />

tanks, and pipeline ruptures. Before 1980, underground gasoline storage tanks<br />

were made from steel; almost half of them were sufficiently corroded to leak<br />

by the time they were 15 years old. Once they descend to groundwater, the<br />

water-soluble components of the gasoline are preferentially leached into the<br />

water and can migrate rapidly in the dissolved state. The BTX component,<br />

which is the most soluble of the hydrocarbons, often occurs at concentrations<br />

of 1–50 ppb in groundwater. However, the alkylated benzenes are rapidly<br />

degraded by aerobic bacteria and consequently are not long-lasting.<br />

The MTBE component of gasoline (Chapter 7) is more water-soluble<br />

than the hydrocarbons, but unlike them, it is not readily biodegraded. It is<br />

not highly toxic. The main problem is the odor and taste that it gives to<br />

water; as little as 15 ppb in water can be tasted or smelled. MTBE contamination<br />

of well water, albeit at low levels, has become of concern in the United<br />

States since it has occurred at about a quarter of a million sites.<br />

The Ultimate Sink for Organic Contaminants<br />

in Groundwater<br />

The subsequent behavior of the organic compounds that do migrate to the<br />

water table depends significantly upon their density relative to that of 1.0 g/mL<br />

of water. Liquids that are less dense (“lighter”) than water and have low solubility<br />

in it form a mass that floats on the top of the water table. All hydrocarbons<br />

having a small or medium molecular mass belong to this group,<br />

including the BTX fraction of gasolines and other petroleum products (see<br />

Table 14-1). In contrast, polychlorinated solvents are more dense (“heavier”)<br />

than water and insoluble in it, so they tend to sink deeply into aquifers;<br />

important examples are methylene chloride, chloroform, carbon tetrachloride,<br />

1,1,1-trichloroethane, TCE, and PCE (see Table 14-1). Nonchlorinated but<br />

insoluble high-molecular-weight organic materials, such as creosote and coal<br />

tar, also belong to the heavier-than-water group. These substances are sometimes<br />

referred to as dense nonaqueous-phase liquids, DNAPLs.<br />

Although the oily liquid blobs that these organic compounds form generally<br />

are found in an aquifer at a position either directly below their original<br />

point of entry into the soil or close to it, the implication that they are horizontally<br />

immobile is misleading. Very slowly—in a process that often takes<br />

decades or centuries to complete—these low-solubility compounds gradually<br />

dissolve in the water that passes over the blob and so provide a continuous<br />

supply of contaminants to the groundwater. The complete removal of such<br />

deposits usually is not feasible since they may exist as several blobs whose<br />

exact location is difficult to pinpoint. In addition, disturbance of the deposit<br />

during removal or treatment may increase its net exposure to the water phase.<br />

Even removing 90% of the substance does not necessarily result in reduction<br />

109


630 Chapter 14 The Pollution and Purification of Water<br />

FIGURE 14-6 The contamination<br />

of groundwater<br />

by organic chemicals.<br />

Aquifer<br />

Direction of<br />

groundwater flow<br />

Surface source<br />

BTX<br />

Chlorinated<br />

solvents<br />

Water table<br />

Plume of<br />

BTX-polluted<br />

water<br />

Plume<br />

of solventpolluted<br />

water<br />

of its groundwater concentration. Thus, plumes of polluted water grow, in the<br />

direction of the water’s flow, and thereby contaminate the bulk of the aquifer<br />

(see Figure 14-6). Because of such contamination, many wells used for drinking<br />

water have had to be closed.<br />

Decontamination of Groundwater:<br />

Physical and Chemical Processes<br />

In the last two decades, considerable energy and money have been spent in<br />

the United States on attempts to control aquifer pollution by the oily liquids<br />

discussed above. Dense organic leachates, especially PCE and TCE, have<br />

contaminated the groundwater that lies below the waste sites associated with<br />

the U.S. Superfund remediation initiative (see Chapter 16). Unfortunately,<br />

no easy cure to the problem of contamination has been found. Control usually<br />

consists of pump-and-treat systems that pump contaminated water from the<br />

aquifer, treat it to remove its organic contaminants (using methods of the type<br />

described later in this chapter), and return the cleaned water to the aquifer or<br />

to some other water body. Alternatively, a fine mist of the contaminated<br />

groundwater is sprayed into the air above agricultural land using a long, moveable<br />

sprinkler system; the contaminant volatile organic compounds (VOCs)<br />

evaporate into the air and the cleansed water is used for irrigation.<br />

110


Groundwater: Its Supply, Chemical Contamination, and Remediation 631<br />

The volume of water that must be pumped and treated in a given<br />

aquifer is huge. For organic contaminants with low water solubility, recontamination<br />

of water returned to the aquifer by additional dissolution from<br />

the blob will occur. Consequently, the treatment systems must operate in<br />

perpetuity, and there are already thousands of them spread across the<br />

United States.<br />

Both in situ heating, to vaporize the organic liquids so their vapors rise<br />

to the soil surface, and the addition of oxidants to convert the substances to<br />

products such as carbon dioxide have been tried in some locations. Typically,<br />

temperatures close to 100°C are used in heating, though it is not known if<br />

this is optimal in most cases. Heating and/or the production of gases or precipitates<br />

by oxidation may inadvertently change the geologic and biological<br />

conditions in the immediate vicinity of the treatment, with unforeseen<br />

effects on the distribution and mobility of the remaining pollutant.<br />

Decontamination of Groundwater:<br />

Bioremediation and Natural Attenuation<br />

Bioremediation is the term applied to the decontamination of water or soil<br />

using biochemical rather than chemical or physical processes. Recently there<br />

has been interesting progress reported in using bioremediation to cleanse<br />

water of chlorinated ethene solvent contamination.<br />

The biodegradation of chloroethenes by aerobic bacteria becomes less<br />

and less efficient as the extent of chlorination increases, so it is ineffective<br />

for perchloroethene. However, under anaerobic conditions, the reductive<br />

biodegradation of PCE and TCE proceeds more quickly, particularly if an easily<br />

oxidized substance such as methanol is added to supply electrons for the<br />

reduction processes. Unfortunately, the stepwise dechlorination of these<br />

compounds proceeds through vinyl chloride, CH2" CHCl, a known carcinogen.<br />

Recently, a bacterium has been discovered that removes all the<br />

chlorine from organic solvents such as TCE and PCE.<br />

Owing to the high cost and limited effectiveness of many groundwater<br />

cleanup technologies, the inexpensive process of natural attenuation—<br />

allowing natural biological, chemical, and physical processes to treat groundwater<br />

contaminants—has become popular. Indeed, it is now used at more<br />

than 25% of Superfund program sites in the United States and is the leading<br />

method to remedy the contamination of groundwater from leaking underground<br />

storage sites.<br />

However, there is great controversy about whether or not natural<br />

attenuation is an appropriate strategy for managing groundwater contamination:<br />

Many environmentalists feel that it is a cheap way for industry to<br />

avoid expensive cleanup costs. The U.S. National Research Council in<br />

1997 appointed a committee to determine which pollutants could be<br />

treated successfully by this technique. Table 14-2 summarizes their results.<br />

111


632 Chapter 14 The Pollution and Purification of Water<br />

TABLE 14-2<br />

Likelihood of Success of Groundwater Remediation<br />

by Natural Attenuation for Various Substances<br />

Dominant Attenuation Likelihood of Success Given<br />

Chemical Class Processes Current Level of Understanding<br />

Organic Compounds<br />

Hydrocarbons<br />

BTEX Biotransformation High<br />

Gasoline, fuel oil Biotransformation Moderate<br />

Nonvolatile aliphatic Biotransformation,<br />

compounds immobilization Low<br />

PAHs Biotransformation,<br />

immobilization Low<br />

Creosote Biotransformation,<br />

Oxygenated hydrocarbons<br />

immobilization Low<br />

Low-molecular-weight<br />

alcohols, ketones, esters<br />

Biotransformation High<br />

MTBE<br />

Halogenated aliphatics<br />

Biotransformation Low<br />

PCE, TCE, carbon tetrachloride Biotransformation Low<br />

TCA Biotransformation,<br />

abiotic transformation Low<br />

Methylene chloride Biotransformation High<br />

Vinyl chloride Biotransformation Low<br />

Dichloroethylene<br />

Halogenated aromatics<br />

Highly chlorinated<br />

Biotransformation Low<br />

PCBs, tetrachlorodibenzofuran, Biotransformation,<br />

pentachlorophenol,<br />

multichlorinated benzenes<br />

Less chlorinated<br />

immobilization Low<br />

PCBs, dioxins Biotransformation Low<br />

Monochlorobenzene Biotransformation Moderate<br />

Inorganic Substances<br />

Metals<br />

Ni Immobilization Moderate<br />

Cu, Zn Immobilization Moderate<br />

Cd Immobilization Low<br />

Pb Immobilization Moderate<br />

112


TABLE 14-2<br />

Groundwater: Its Supply, Chemical Contamination, and Remediation 633<br />

Likelihood of Success of Groundwater Remediation<br />

by Natural Attenuation for Various Substances<br />

Dominant Attenuation Likelihood of Success Given<br />

Chemical Class Processes Current Level of Understanding<br />

Cr Biotransformation,<br />

immobilization Low to moderate<br />

Hg Biotransformation,<br />

immobilization Low<br />

Nonmetals<br />

As Biotransformation,<br />

immobilization Low<br />

Se Biotransformation,<br />

immobilization Low<br />

Oxyanions<br />

Nitrate Biotransformation Moderate<br />

Perchlorate Biotransformation Low<br />

Source: Adapted from J. A. Macdonald, “Evaluating Natural Attenuation for Groundwater Cleanup,” Environmental Science and<br />

Technology (1 August 2000): 346A.<br />

Only three pollutants are highly likely to be successfully treated by natural<br />

attenuation:<br />

• BTEX hydrocarbons (i.e., BTX hydrocarbons plus ethylbenzene),<br />

• low-molecular-weight oxygen-containing organics, and<br />

• methylene chloride.<br />

In all three cases, biotransformation is the dominant process by which attenuation<br />

occurs. Notice that neither MTBE nor highly chlorinated organics,<br />

including TCE and PCE, are usually successfully treated in this way, nor is<br />

mercury or perchlorate ion.<br />

Decontamination of Groundwater: In Situ Remediation<br />

Scientists have developed a promising in situ technique for treating groundwater<br />

contaminated by volatile (mainly C1 and C2 ) chlorinated organics.<br />

They construct an underground permeable “wall” of material (mostly coarse<br />

sand) along the path of the water. The water is cleansed as a result of its<br />

passage through the wall and never has to be pumped out of the ground (see<br />

Figure 14-7).<br />

The ingredient that is placed within the sand bed and that chemically<br />

cleans the water is metallic iron, Fe 0 , in the form of small granules, a common<br />

113


634 Chapter 14 The Pollution and Purification of Water<br />

FIGURE 14-7 In situ<br />

purification of groundwater<br />

using an “iron wall.”<br />

Permeable subsurface<br />

treatment wall composed<br />

of iron filings and sand<br />

Clean<br />

groundwater<br />

Gravel<br />

Stream of<br />

chemicals<br />

waste product of manufacturing processes. When placed in contact with certain<br />

chlorinated organics dissolved in water, the iron acts as a reducing agent,<br />

giving up electrons to form the ferrous, or Fe(II), ion Fe 2 , which dissolves<br />

in the water:<br />

Fe(s) 9: Fe 2 (aq) 2 e <br />

Usually, these electrons are donated to chloroorganic molecules that are temporarily<br />

adsorbed onto the metal’s surface; the chlorine atoms contained in<br />

these molecules are consequently reduced to chloride ions, Cl , which are<br />

released into aqueous solution. This technique is an example of reductive<br />

degradation. For example, the reduction of trichloroethene to its completely<br />

dechlorinated form, ethene, can be written in unbalanced form as<br />

C 2 HCl 3 9: C 2 H 4 3 Cl <br />

Chemical source<br />

Direction of<br />

groundwater<br />

flow<br />

Upon application of a standard redox balancing technique for alkaline solution,<br />

we obtain the balanced half-reaction<br />

C 2 HCl 3 3 H 2 O 6 e 9: C 2 H 4 3 Cl 3 OH <br />

Combination of the half-reactions (after tripling the oxidation step to ensure<br />

the number of electrons lost and gained is the same) yields the overall reaction<br />

3 Fe(s) C 2HCl 3 3 H 2O 9: 3 Fe 2 (aq) C 2H 4 3 Cl 3 OH <br />

One of the by-products of the reaction is hydroxide ion, OH . Recall<br />

from Chapter 13 that in limestone-rich areas, the groundwater contains<br />

114


significant concentrations of dissolved calcium bicarbonate, Ca(HCO 3 ) 2 .<br />

The hydroxide ions produced in the groundwater remediation reaction react<br />

with bicarbonate to produce carbonate ion, CO 3 2 , which combines with dissolved<br />

calcium ions to produce insoluble calcium carbonate, CaCO 3, which<br />

then precipitates in the sand–iron mixture.<br />

Field trials indicate that this new technology can work successfully for<br />

several years at least, and it may replace the pump-and-treat methods for<br />

many applications involving chlorinated methanes and ethanes dissolved in<br />

underground water.<br />

Recently, it has been found that coating the iron filings with nickel<br />

speeds up the rate of degradation of the organic compounds by a factor of 10;<br />

with this modification, the technique may be even more useful than first<br />

imagined. In addition, it has been discovered that the elemental iron in the<br />

barriers will reduce soluble Cr 6 ions to insoluble Cr 3 oxides and can therefore<br />

remediate groundwater contaminated by Cr 6 . (The environmental chemistry<br />

of chromium is discussed in more detail in Chapter 15). A technique for<br />

the in situ creation of elemental iron from its ions (Fe 2 and Fe 3 ) by the<br />

injection of aqueous reducing agents has also been tested.<br />

An in situ technique of treating TCE and PCE by hydrogenation has<br />

been developed. The process uses dissolved H 2 gas to rapidly dechlorinate<br />

these two organics, eventually forming ethane and HCl. The reaction, which<br />

uses a palladium catalyst, can be done within a well bore so that the water<br />

need not be brought up to the surface.<br />

PROBLEM 14-7<br />

The dissolution of iron in the process described above produces some molecular<br />

hydrogen gas, H 2 . Show by balanced equation(s) how the hydrogen could<br />

arise from the reduction of water rather than of TCE.<br />

PROBLEM 14-8<br />

Suppose that this “iron wall” technology reduced an appreciable fraction<br />

of TCE to vinyl chloride rather than completely to ethene. Why would this<br />

be an unacceptable result environmentally? (Note that in practice a sufficiently<br />

thick wall of iron is used to convert any vinyl chloride by-product to<br />

ethene.)<br />

PROBLEM 14-9<br />

Groundwater: Its Supply, Chemical Contamination, and Remediation 635<br />

Deduce the overall reaction by which perchloroethene is converted to<br />

ethene by metallic iron.<br />

115


636 Chapter 14 The Pollution and Purification of Water<br />

PROBLEM 14-10<br />

At one test site for this remediation process, the water contained 270 ppm<br />

TCE and 53 ppm perchloroethene. Calculate the mass of iron required to<br />

remediate 1 L of this groundwater.<br />

The Chemical Contamination and<br />

Treatment of Wastewater and Sewage<br />

Most municipalities treat the raw sewage collected from homes, buildings,<br />

and industries (including food processing plants) through a sanitary sewer<br />

system before the liquid residue is deposited into a nearby source of natural<br />

waters, whether a river, lake, or ocean. In contrast, since the rainwater and<br />

melted snow that drain from streets and other paved surfaces are usually not<br />

highly contaminated, they are often collected separately by storm sewers and<br />

deposited directly into a body of natural water. Unfortunately, in some<br />

municipalities, storm-driven overflow occurs in the sanitary sewer system<br />

and the overflow is combined with storm water and deposited, untreated,<br />

into waterways.<br />

The main component of sewage—other than water—is organic matter of<br />

biological origin. It occurs mainly as particles—ranging from those of macroscopic<br />

size large enough to be trapped (together with such objets d’art as facial<br />

tissues, stones, socks, tree branches, condoms, and tampon applicators) by<br />

mesh screens to those which are microscopic in size and are suspended in the<br />

water as large colloids.<br />

Sewage Treatment<br />

In the primary (or mechanical) treatment stage of wastewater (see the<br />

schematic diagram in Figure 14-8), the larger particles—including sand and<br />

silt—are removed by allowing the water to flow across screens and slowly<br />

along a lagoon or settling basin. A sludge of insoluble particles forms at the<br />

bottom of the lagoon, while “liquid grease” (a term which here includes not<br />

only fat, oils, and waxes but also the products formed by the reaction of soap<br />

with calcium and magnesium ions) forms a lighter-than-water layer at the top<br />

and is skimmed off. About 30% of the biochemical oxygen demand (BOD,<br />

Chapter 13) of the wastewater is removed by the primary treatment process,<br />

even though this stage of the process is entirely mechanical in nature. The<br />

treatment and disposal of the sludge is discussed later.<br />

After passing through conventional primary treatment, the sewage water<br />

has been much clarified but still has a very high BOD—typically several<br />

hundred milligrams per liter (ppm)—and is detrimental to fish life if released<br />

at this stage (as occurs in some jurisdictions that discharge into the ocean).<br />

The high BOD is due mainly to organic colloidal particles. In the secondary<br />

(biological) treatment stage, most of this suspended organic matter as well as<br />

116


Wastewater<br />

Water<br />

Grease<br />

Solid matter<br />

The Chemical Contamination and Treatment of Wastewater and Sewage 637<br />

Primary treatment Secondary treatment Tertiary<br />

Microorganism-<br />

treatment<br />

Grease<br />

catalyzed oxidation<br />

Sludge<br />

Disposal<br />

that actually dissolved in the water is biologically oxidized by microorganisms<br />

to carbon dioxide and water or converted to additional sludge, which can<br />

readily be removed from the water. Either the water is sprinkled onto a bed of<br />

sand and gravel or of plastic covered with the aerobic bacteria (trickling filters),<br />

or it is agitated in an aeration reactor (activated sludge process) in order to<br />

effect the microorganism-driven reaction. The system is kept well-aerated to<br />

speed the oxidation. In essence, by deliberately maintaining in the system a<br />

high concentration of aerobic microorganisms, especially bacteria, the same<br />

biological degradation processes that would require weeks to occur in open<br />

waters are accomplished in several hours.<br />

The biological oxidation processes of secondary treatment reduce the<br />

BOD of the polluted water to less than 100 ppm, which is about 10% of the<br />

original concentration in the untreated sewage. For comparison, Canadian<br />

wastewater quality standards require that the BOD in treated water be 20 ppm<br />

or less. The process of nitrification also occurs to some extent, converting<br />

organic nitrogen compounds to nitrate ion and carbon dioxide (see below). In<br />

summary, the secondary treatment of wastewater involves biochemical reactions<br />

that oxidize much of the oxidizable organic material that was not removed in<br />

the first stage. Treated water diluted with a greater amount of natural water<br />

can support aquatic life. Normally, municipalities take the water produced by<br />

secondary treatment and disinfect it by chlorination or irradiation with UV<br />

light before pumping it into a local waterway. Recent research in Japan has<br />

shown that chlorination of the effluent before its release produces mutagenic<br />

compounds, presumably by interaction of chlorine-containing substances with<br />

the organic matter that remains in the water.<br />

A few municipalities employ tertiary (or advanced or chemical) wastewater<br />

treatment as well as primary and secondary ones. In the tertiary phase,<br />

specific substances are removed from the partially purified water before its<br />

final disinfection. In some cases, the water produced by tertiary treatment is<br />

117<br />

Sludge<br />

Removal<br />

of various<br />

chemicals<br />

River<br />

or lake<br />

FIGURE 14-8 The<br />

common stages in the<br />

treatment of wastewater.


638 Chapter 14 The Pollution and Purification of Water<br />

of sufficiently high quality to use as drinking water. Alternatively, river<br />

water into which the effluent from sewage treatment plants has been<br />

deposited is used by municipalities downstream as drinking water. The reuse<br />

of water after it has been cleansed is particularly prevalent in Europe, where<br />

less fresh water is available than in North America and the population<br />

density is high.<br />

Depending upon locale, tertiary treatment can include some or all of the<br />

following chemical processes:<br />

• further reduction of BOD by removal of most remaining colloidal material,<br />

using an aluminum salt in a process that operates in the same manner as<br />

described previously for the purification of drinking water;<br />

• removal of dissolved organic compounds (including chloroform) and<br />

some heavy metals by their adsorption onto activated carbon, over which the<br />

water is allowed to flow (see Box 14-1);<br />

• phosphate removal (as discussed in the next section; some phosphorus is<br />

removed in the secondary treatment stage since microbes incorporate it as a<br />

nutrient for their growth);<br />

• heavy metal removal by the addition of hydroxide or sulfide ions, S 2 , to<br />

form insoluble metal hydroxides or sulfides, respectively (see Chapter 15);<br />

• iron removal by aeration at a high pH to oxidize it to its insoluble Fe 3<br />

state, possibly in combination with use of a strong oxidizing agent to destroy<br />

organic ligands bound strongly to the Fe 2 ion, which would otherwise prevent<br />

its oxidation; and<br />

• removal of excess inorganic ions, as discussed below.<br />

In some wastewaters, the further removal of nitrogen compounds—usually<br />

either ammonia or organic nitrogen compounds—is deemed necessary. Ammonia<br />

removal can be achieved by raising the pH to about 11 (with lime) to convert<br />

most ammonium ion to its molecular form, ammonia, NH3 , followed by<br />

bubbling air through the water to air-strip it of its dissolved ammonia gas. This<br />

process is relatively expensive, however, because it is energy-intensive. Ammonium<br />

ion can also be removed by ion exchange, using certain resins that have<br />

their exchange sites initially populated by sodium or calcium ions.<br />

Alternatively, both organic nitrogen and ammonia can be removed by<br />

first using nitrifying bacteria to oxidize all the nitrogen to nitrate ion. Then<br />

the nitrate is subjected to denitrification by bacteria to produced molecular<br />

nitrogen, N2, which bubbles out of the water. Since this reduction step<br />

requires a substance to be oxidized, methanol, CH3OH, is added if necessary<br />

to the water and is converted to carbon dioxide in the process:<br />

bacteria<br />

5 CH 3 OH 6 NO 3 6 H 9: 5 CO2 3 N 2 13 H 2 O<br />

118


Of course, water contaminated by nitrate ion can also be treated by this latter<br />

step. A mathematical analysis of the kinetics of the transformations is<br />

given in Box 14-4.<br />

PROBLEM 14-11<br />

The Chemical Contamination and Treatment of Wastewater and Sewage 639<br />

BOX 14-4 Time Dependence of Concentrations<br />

in the Two-Step Oxidation of Ammonia<br />

The bacteria-catalyzed oxidation of ammonia<br />

(or of other reduced organic nitrogen<br />

compounds) to nitrate is a reaction with two<br />

<br />

main steps, with nitrite ion, NO2 , an intermediate:<br />

Step 1<br />

3<br />

NH3 2 O2 9:<br />

<br />

NO2 H H2O<br />

Step 2<br />

1<br />

<br />

NO2 O2 9: NO3 2<br />

If sufficient oxygen is available, the rate of<br />

each reaction is first-order only in the concentration<br />

of the nitrogen reactant, so the<br />

sequence can be represented as<br />

k 1<br />

k 2<br />

A 9: B 9: C<br />

where A stands for ammonia, B for nitrite ion,<br />

and C for nitrate ion, and k 1 and k 2 are the<br />

pseudo-first-order rate constants. Since the rate<br />

of step 1 depends on the first power of the ammonia<br />

concentration, then the rate of disappearance<br />

of this species is<br />

d[A]<br />

dt k 1[A]<br />

Since B (nitrite) is produced at this rate by<br />

step 1, but is consumed in step 2 by a process<br />

whose rate is proportional to the first power of<br />

its concentration, we can write<br />

These differential equations can be coupled<br />

and integrated to yield the following<br />

expressions for the evolution of [A] and [B]<br />

with time, relative to [A] 0, the original concentration<br />

of A:<br />

[A]/[A] 0 e k 1t<br />

[B]/[A] 0 k 1(e k 1t e k 2t )/(k2 k 1)<br />

As can be seen from the solution to Problem 1,<br />

the concentration of B (nitrite) rises exponentially<br />

at first, reaches a peak value, then<br />

declines slowly. Thus significant concentrations<br />

of nitrite ion occur in water undergoing<br />

the two-step conversion of ammonia to<br />

nitrate.<br />

PROBLEM 1<br />

d[B]<br />

dt k 1[A] k 2[B]<br />

(a) Derive a general expression relating the<br />

time at which [B] reaches a peak to k 1 and k 2 .<br />

(b) Draw a graph showing the evolution of<br />

[A] and of [B] with time for the values k 1 1<br />

and k 2 2.<br />

Given that the K b value for ammonia is 1.8 10 5 , deduce a formula giving<br />

the ratio of ammonia to ammonium ion as a function of the pH of water.<br />

What is the value of this ratio at pH values of 5, 7, 9, and 11?<br />

119


640 Chapter 14 The Pollution and Purification of Water<br />

The Origin and Removal of Excess Phosphate<br />

One of the world’s most famous cases of water pollution involves Lake Erie,<br />

which in the 1960s was said to be dying. Indeed, one of the authors of this<br />

book can recall visiting a once-popular beach on Lake Erie’s north shore in the<br />

early 1970s and being repulsed by the sight and smell of dead, rotting fish on<br />

the shoreline. Lake Erie’s problems stemmed primarily from an excess input of<br />

phosphate ion, PO 4 3 , in the waters of its tributaries. The phosphate sources<br />

were the polyphosphates in detergents (as explained in detail later), raw<br />

sewage, and the runoff from farms that used phosphate products. Since there is<br />

commonly an excess of other dissolved nutrients in lakes, phosphate ion usually<br />

functions as the limiting (or controlling) nutrient for algal growth: The<br />

larger the supply of the ion, the more abundant the growth of algae—and its<br />

growth can be quite abundant indeed. When the vast mass of excess algae<br />

eventually dies and starts to decompose by oxidation, the water becomes<br />

depleted of dissolved oxygen, with the result that fish life is adversely affected.<br />

The lake water also becomes foul-tasting, green, and slimy, and masses of dead<br />

fish and aquatic weeds rot on the beaches. The series of changes, including<br />

rapid degradation and aging, that occur when lakes receive excess plant nutrients<br />

from their surroundings is called eutrophication. When the enrichment<br />

arises from human activities, it is called cultural eutrophication.<br />

To correct the regional problem, the United States and Canada in 1972<br />

signed the Great Lakes Water Quality Agreement. Since that time, over $8 billion<br />

has been spent in building sewage treatment plants to remove phosphates<br />

from wastewater before it reaches the tributaries and the lake itself. In addition,<br />

the levels of polyphosphates in laundry detergents were restricted in Ontario<br />

and in many of the states that border the Great Lakes. The total amount of<br />

phosphorus entering Lake Erie has now decreased by more than two-thirds. As<br />

a result, Lake Erie has sprung back to life: Its once-fouled beaches are regaining<br />

popularity with tourists, and its commercial fisheries have been revived.<br />

As we have pointed out, the presence of excess phosphate ion in natural<br />

waters can have a devastating effect on an aquatic ecology because it overfertilizes<br />

plant life. Formerly, one of the largest sources of phosphate as a pollutant<br />

was detergents, and in the material that follows, the role of such<br />

phosphates is discussed.<br />

The reaction of synthetic detergents with calcium and magnesium ions,<br />

forming complex ions, diminishes the cleansing potential of the detergent.<br />

Polyphosphate ions, which are anions containing several phosphate units<br />

linked by shared oxygens, are added to detergents as builders that preferentially<br />

form soluble complexes with these metal ions and thereby allow the<br />

molecules of the detergent to operate as cleansing agents rather than being<br />

complexed with the Ca 2 and Mg 2 naturally present in the water. Another<br />

role of the builder is to make the wash water somewhat alkaline, which helps<br />

remove the dirt from certain fabrics. With soap itself, the ions form insoluble<br />

complexes that foul the cleaning water.<br />

120


(a)<br />

The Chemical Contamination and Treatment of Wastewater and Sewage 641<br />

O <br />

O9P9O9P9O9P9O <br />

O<br />

O <br />

O<br />

O <br />

O<br />

Great quantities of sodium tripolyphosphate (STP), Na5P3O10, were<br />

formerly added as the builder in most synthetic detergent formulations. As<br />

shown in Figure 14-9a, STP contains a chain of alternating phosphorus and<br />

oxygen atoms, with one or two additional oxygens attached to each phosphorus.<br />

In solution, one tripolyphosphate ion can form a complex with one calcium<br />

ion by forming interactions between three of its oxygen atoms and the<br />

metal ion (Figure 14-9b).<br />

Substances like STP, which have more than one site of attachment to the<br />

metal ion and thereby produce ring structures that each incorporate the<br />

metal, are called chelating agents (from the Greek word for “claw”). Because<br />

several bonds are formed, the resulting chelates are very stable and do not<br />

normally release their metal ions back into the free form. The use of chelating<br />

agents to remove metals from the human body is discussed in Chapter 15.<br />

5<br />

Tripolyphosphate ion, P3O10 , like phosphate ion itself, is a weak base<br />

in aqueous solution and thus provides the alkaline environment that is<br />

required for effective cleaning:<br />

5<br />

P3O10 H2O 9: P3O10H 4 OH <br />

Unfortunately, when wash water containing STP is discarded, the excess<br />

tripolyphosphate enters waterways, where it slowly reacts with water and is<br />

transformed into phosphate ion (sometimes called orthophosphate):<br />

5 3 <br />

P3O10 2 H2O 9: 3 PO4 4 H<br />

Note that when tripolyphosphate decomposes, STP behaves as an acid rather<br />

than a base (since H is formed in the reaction).<br />

Because of environmental concerns, polyphosphates are now used only<br />

sparingly as builders in detergents in many areas of the world. In Canada and<br />

parts of Europe, STP was replaced largely by sodium nitrilotriacetate (NTA)<br />

(see Figure 14-10a). The anion of NTA acts in a similar fashion to that of<br />

STP, chelating calcium and magnesium ions using three of its oxygen atoms<br />

and the nitrogen atom (Figure 14-10b). NTA is not used as a builder in the<br />

United States because of concerns that its slow rate of degradation might lead<br />

to health hazards in drinking water. However, the early experiments with test<br />

animals that led to this concern are open to question, as are fears about its<br />

persistence and its tendency to solubilize heavy metals into water supplies.<br />

Other builders now used include sodium citrate, sodium carbonate (washing<br />

soda), and sodium silicate. Currently, substances called zeolites are also<br />

employed as detergent builders. Zeolites are abundant aluminosilicate minerals<br />

(b)<br />

121<br />

O<br />

O O O<br />

P<br />

Ca2 O P<br />

O9P<br />

O<br />

O <br />

O<br />

O <br />

FIGURE 14-9 Structure<br />

of the polyphosphate ion:<br />

(a) uncomplexed and<br />

(b) complexed with<br />

calcium ion.


642 Chapter 14 The Pollution and Purification of Water<br />

FIGURE 14-10 Structure<br />

of the nitrilotriacetate ion:<br />

(a) uncomplexed and<br />

(b) complexed with<br />

calcium ion.<br />

O<br />

CH 2 9C<br />

(a)<br />

N<br />

(b)<br />

O<br />

C<br />

CH2 O O<br />

CH29C (see Chapter 16) consisting of sodium, aluminum, silicon, and oxygen. The<br />

latter three elements are bonded together to form cages, which the sodium ions<br />

can enter. In the presence of calcium ion, zeolites exchange their sodium ions<br />

for Ca 2 (though not for Mg 2 ), thereby sequestering it in a manner similar to<br />

polyphosphates. Like polyphosphates, they also control pH. One disadvantage<br />

to the use of zeolites is that they are insoluble, so their use increases the amount<br />

of sludge that must be removed at wastewater treatment plants.<br />

Phosphate ion can be removed from municipal and industrial wastewater<br />

by the addition of sufficient calcium as the hydroxide Ca(OH) 2 , so<br />

that insoluble calcium phosphates such as Ca 3 (PO 4 ) 2 and Ca 5 (PO 4 ) 3 OH<br />

are formed as precipitates that can then be readily removed. Phosphate<br />

removal could be a standard practice in the treatment of wastewater, but it<br />

is not yet practiced in all cities. Some policymakers believe that the optimum<br />

environmental solution is to use polyphosphates, rather than some<br />

other builder, in detergents and then to efficiently remove phosphates at<br />

wastewater treatment plants.<br />

Geographically, phosphate ion enters waterways from both point and nonpoint<br />

sources. Point sources are specific locations such as factories, landfills,<br />

and sewage treatment plants that discharge pollutants. Nonpoint sources are<br />

numerous large land areas such as farms, logged forests, septic tanks, golf courses<br />

and individual domestic lawns, stormwater runoff, and atmospheric deposition.<br />

Although each nonpoint source may provide a small amount of pollution, on<br />

account of the large number of them involved they can generate larger total<br />

quantities than do point sources. For example, now that sewage treatment<br />

plants and detergent controls have been instituted, much of the remaining<br />

phosphate arises from nonpoint agricultural sources in many areas.<br />

Green Chemistry: Sodium Iminodisuccinate—<br />

A Biodegradable Chelating Agent<br />

Because most chelating agents are not biodegradable or are only slowly<br />

biodegradable, not only do they place a load on the environment (e.g., phosphates<br />

acts as nutrients), but it may be necessary to remove them during<br />

122<br />

O<br />

O <br />

H 2 C<br />

H 2 C<br />

CH 2<br />

O<br />

C<br />

C<br />

C<br />

O<br />

O<br />

O <br />

N Ca 2<br />

O O


The Chemical Contamination and Treatment of Wastewater and Sewage 643<br />

treatment of wastewater in a wastewater<br />

treatment plant. Unlike many chelating<br />

agents, sodium iminodisuccinate (IDS, see<br />

Figure 14-11) [also known as D,L-aspartic-N-<br />

(1,2-dicarboxylethyl) tetrasodium salt] readily<br />

degrades in the environment. Not only is IDS<br />

biodegradable, it is also nontoxic.<br />

IDS can be used as an effective chelating<br />

agent for absorption of agricultural nutrients,<br />

metal ion scavenging in photographic processing, groundwater remediation,<br />

and as a builder in detergents and household and industrial cleaners. The<br />

Bayer Corporation won a Presidential Green Chemistry Challenge Award in<br />

2001 for the development of IDS as a chelating agent and for its synthesis<br />

from maleic anhydride (Figure 14-11). This synthesis is accomplished under<br />

mild conditions, in water as the only solvent. The excess ammonia is recycled<br />

back into the production of more IDS. This synthesis stands in stark contrast<br />

to typical syntheses of aminocarboxylate chelating agents, which employ<br />

hydrogen cyanide as a reagent. Bayer markets IDS as a chelating agent under<br />

the name Baypure.<br />

Reducing the Salt Concentration in Water<br />

The decomposition of organic and biological substances during the secondary<br />

phase of wastewater treatment usually results in the production of inorganic<br />

salts, many of which remain in the water even after the techniques listed<br />

above have been applied. Water can also become salty due to its use in irrigation<br />

or because water softener units have been recharged and their discharge<br />

disposed of as sewage. Inorganic ions can be removed from water by desalination<br />

by using one of the techniques listed below or using the precipitation methods<br />

mentioned above.<br />

• Reverse osmosis As previously mentioned, this technique is also used to<br />

produce drinking water from salty water, such as seawater.<br />

• Electrodialysis Here a series of membranes permeable either only to<br />

small inorganic cations or only to small inorganic anions are set up vertically<br />

in an alternating fashion (see Figure 14-12) within an electrochemical cell.<br />

Direct current is applied across the water, so cations migrate toward the cathode<br />

and anions toward the anode. The liquid in alternating zones becomes<br />

more concentrated (enriched) or less concentrated (purified) in ions; eventually<br />

the ion-concentrated water can be disposed of as brine and the purified<br />

water released into the environment. This technology is also used to desalinate<br />

seawater for drinking purposes.<br />

• Ion exchange Some polymeric solids contain sites that hold ions relatively<br />

weakly, so one type of ion can be exchanged for another of the same charge that<br />

123<br />

O<br />

O<br />

O<br />

maleic anhydride<br />

NH 3<br />

NaOH<br />

NaO<br />

NaO<br />

O<br />

NH<br />

O<br />

O O<br />

sodium iminodisuccinate<br />

ONa<br />

ONa<br />

FIGURE 14-11 Synthesis<br />

and structure of sodium<br />

iminodisuccinate, a<br />

biodegradable chelating<br />

agent.


644 Chapter 14 The Pollution and Purification of Water<br />

Outgoing<br />

brine<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

+<br />

+<br />

+<br />

–<br />

+<br />

+<br />

+<br />

–<br />

–<br />

–<br />

+<br />

–<br />

–<br />

–<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

Cathode Anode<br />

Incoming<br />

salty water<br />

+ Cations<br />

– Anions<br />

FIGURE 14-12 Electrodialysis<br />

unit (schematic) for<br />

the desalination of water.<br />

[Adapted from S. E. Manahan.<br />

1994. Environmental Chemistry,<br />

6th ed. Boca Raton, FL:<br />

Lewis Publishers.]<br />

Outgoing<br />

purified<br />

water<br />

Cationpermeable<br />

membrane<br />

Anionpermeable<br />

membrane<br />

Outgoing<br />

brine<br />

PROBLEM 14-12<br />

happens to pass by it. Ion exchange resins<br />

can be formulated to possess either cationic<br />

or anionic sites that function in this manner.<br />

The exchange sites of a cationic resin of this<br />

type are initially occupied by H ions, and<br />

the exchange sites of an anionic exchange<br />

resin are occupied by OH ions. When water<br />

polluted by M and X ions is passed sequentially<br />

through the two resins, the H ions on<br />

the first are replaced by M , and then the<br />

OH ions on the second resin are replaced by<br />

X . Thus the water that has passed through<br />

contains H and OH ions, rather than<br />

those of the salt, which remain behind in<br />

the resins. Of course, these two ions immediately<br />

combine to form more water molecules.<br />

Thus ion exchange can be used to<br />

remove salts, including those of heavy<br />

metals, from wastewater.<br />

Water polluted by inorganic ions could be purified by distilling it or by freezing<br />

it. Why do you think such techniques are not generally used on a mass<br />

scale to purify water?<br />

Transition metal cations can be removed from water using either precipitation<br />

or reduction techniques, in either case to form insoluble solids. Precipitation<br />

of sulfides or hydroxides has already been mentioned; a disadvantage<br />

of the latter is the production of a voluminous sludge that must be disposed of<br />

in an acceptable manner. Electrolytic reduction of metals leads to their deposition<br />

on the cathode. If, instead of the elemental metal, a concentrated<br />

aqueous solution of it is desired, the deposited metal can be reoxidized chemically<br />

by adding hydrogen peroxide or electrolytically by reversing the polarity<br />

of the cell.<br />

The Biological Treatment of Wastewater and Sewage<br />

An alternative to the processing of wastewater through a conventional treatment<br />

plant in small communities is biological treatment in an artificial marsh<br />

(also called a constructed wetland) that contains plants such as bullrushes and<br />

reeds. The decontamination of the water is accomplished by the bacteria and<br />

other microbes that live among the plants’ roots and rhizomes. The plants<br />

124


The Chemical Contamination and Treatment of Wastewater and Sewage 645<br />

themselves take up metals through their root systems and concentrate<br />

contaminants within their cells. In facilities that have been constructed to<br />

deal with sewage, primary treatment to filter out solids, etc. in a lagoon is<br />

usually implemented before the wastewater is pumped to the marsh, where<br />

the equivalent of secondary and tertiary treatment occurs. The plant growth<br />

uses up the pollutants and increases the pH—which serves to destroy some<br />

harmful microorganisms.<br />

One advantage of biological treatment is that great amounts of sludge are<br />

not generated, in contrast to conventional treatments. Furthermore, it requires<br />

neither the addition of synthetic chemicals nor the input of commercial<br />

energy. Among the problems in such facilities are decaying vegetation, which<br />

must be limited so that the BOD of the processed water does not rise too<br />

much, and the fact that the marshes usually require a great deal of land unless<br />

they are constructed so that part of the routing is vertical.<br />

In many rural and small communities, septic tanks are used to decontaminate<br />

sewage since central sewage facilities are not available. These underground<br />

concrete tanks receive the wastewater, often from only one home.<br />

Although solids settle in the tank, grease and oil rise to the top, from which<br />

they are periodically removed. The bacteria in the wastewater feed on the<br />

bottom sludge, thereby liquefying the waste. Partially purified water flows out<br />

of the tank into an underground drain, where further decontamination takes<br />

place. The system is relatively passive, compared to central facilities, and as<br />

in the case of artificial marshes, time is required for the processes to occur. In<br />

addition, nitrogen compounds are converted to nitrate, but the latter is not<br />

reduced to molecular nitrogen, so groundwater under the septic system can<br />

become contaminated by nitrate, as discussed earlier in this chapter.<br />

Drugs in Wastewater from Sewage Treatment Plants<br />

In recent years, trace concentrations of various drugs—prescription, overthe-counter,<br />

illegal, and veterinary—have been detected in the waters leading<br />

from sewage treatment plants, and in rivers and streams into which this<br />

water then flows, at concentrations up to the ppb level. About 100 substances<br />

have been detected in various rivers, lakes, and coastal waters. The<br />

substances—commonly including estradiol, ibuprofen, the antidepressant drug<br />

Prozac (fluoxetine), the anti-epileptic drug carbamazepine, and degradation<br />

products associated with cholesterol-reducing pharmaceuticals—are present<br />

in raw sewage after their excretion in urine or feces from humans and animals<br />

since most drugs are poorly absorbed and metabolized by the body. They also<br />

result from the disposal of unused or expired medication in toilets.<br />

Most commonly, concentrations of drugs in drinking water are at the<br />

parts-per-trillion level, so their risk to human health is probably small.<br />

Research is under way to determine whether there could be effects on human<br />

health from sustained exposure to a combination of these substances. The<br />

125


646 Chapter 14 The Pollution and Purification of Water<br />

synthetic hormones are thought to pose the greatest risk to aquatic species.<br />

Certain fish have been found to undergo some skewed sexual development<br />

due to exposure to sewage effluent containing the synthetic estrogen in birth<br />

control pills (see Chapter 12).<br />

The Treatment of Cyanides in Wastewater<br />

The cyanide ion, CN , binds strongly to many metals, especially those of<br />

the transition series, and is often used to extract them from mixtures. Consequently,<br />

cyanide is widely used in mining, refining, and electroplating metals<br />

such as gold, cadmium, and nickel. Unfortunately, cyanide ion is very poisonous<br />

to animal life since it binds strongly to metal ions in living matter,<br />

e.g., to the iron in proteins that are necessary for molecular oxygen to be utilized<br />

by cells.<br />

Cyanide is a very stable species and does not quickly decompose on its<br />

own or in the environment. Thus it is an important water pollutant and<br />

should be destroyed chemically rather than simply disposed of in a waterway.<br />

We can deduce the type of treatment that will be effective for cyanide by<br />

considering its acid–base and redox characteristics. Cyanide ion is the conjugate<br />

base of the weak acid hydrocyanic acid, HCN, which has limited solubility<br />

in water. Thus, acidification of cyanide solutions will result in the release of<br />

poisonous HCN gas from it and therefore is not a good solution to the problem<br />

of cyanide contamination.<br />

The redox chemistry of cyanide ion can be predicted by considering the<br />

oxidation numbers of the two atoms involved. If nitrogen, the more electronegative<br />

atom, is considered to be in its fully reduced 3 oxidation form,<br />

then the carbon must be 2. Thus one way to destroy cyanide ion is to oxi-<br />

<br />

dize the carbon more fully, to 4 as it is in CO2 and HCO3 . This oxidation<br />

can be accomplished by dissolved molecular oxygen if high temperatures and<br />

elevated air pressures are used:<br />

2 0<br />

4 2<br />

2 CN O 2 4 H 2 O 2 HCO 3 2 NH3<br />

3<br />

The use of stronger oxidizing agents, such as Cl 2 or ClO , not only oxidizes<br />

the carbon from 2 to 4, but can also oxidize the nitrogen from the 3<br />

state to the zero oxidation number of molecular nitrogen:<br />

2 3 0 4<br />

0<br />

1<br />

2 CN 5 Cl 2 8 OH 2 CO 2 N 2 10 Cl 4 H 2 O<br />

(Four of the ten electrons gained collectively by the chlorines are used by the<br />

carbons and six by the nitrogens in this overall process.) Other oxidizing<br />

agents that are used in cyanide treatment include hydrogen peroxide, H 2 O 2 ,<br />

and/or molecular oxygen, in both cases with a copper salt added as a catalyst.<br />

126


The process can also be carried out electrochemically for high cyanide concentrations;<br />

the remaining low concentration can be subsequently oxidized<br />

by ClO .<br />

Sodium cyanide is now used in some shallow tropical waters such as those<br />

in Indonesia to stun reef fish so that they can be captured and sold live as<br />

seafood or pets. Unfortunately, the cyanide kills smaller fish and destroys the<br />

coral.<br />

PROBLEM 14-13<br />

If an oxidizing agent even more powerful than chlorine or hypochlorite were<br />

to be used in the treatment of cyanide, what other possibilities for the nitrogencontaining<br />

product would there be?<br />

PROBLEM 14-14<br />

The Chemical Contamination and Treatment of Wastewater and Sewage 647<br />

For HCN in water, K a 6.0 10 10 . Calculate the fraction of cyanide that<br />

exists as the anion rather than in the molecular form at pH values of 4, 7,<br />

and 10.<br />

The Disposal of Sewage Sludge<br />

The sludge from both the primary and secondary treatment stages of sewage<br />

is principally water and organic matter. It can be digested anaerobically, in<br />

a process that takes several weeks to complete. Bacteria levels in the sludge<br />

are not thereby completely eliminated, but the levels are reduced about a<br />

thousandfold. The sludge that remains after this further organic decomposition<br />

has occurred and after the supernatant water is removed is sometimes<br />

then incinerated or simply dumped into a landfill or into a water body such<br />

as the ocean. However, sludge is high in plant nutrients, so about half the<br />

sewage sludge in North America and Europe is spread on farm fields, golf<br />

courses, and even residential lawns as low-grade fertilizer sometimes called<br />

biosolid.<br />

Unfortunately, sewage sludge may contain toxic substances, which potentially<br />

could be incorporated into food grown on the land or could contaminate<br />

groundwater under the fields. In particular, heavy metal concentrations often<br />

are higher in sewage sludge than in soil, principally because industrial wastes<br />

are sometimes released directly into sewage lines shared by households. For<br />

example, the lead level in municipal sludge can range from several hundred<br />

to several thousand parts per million, compared to an average of about 10 ppm<br />

in the Earth’s crust. In a few communities, an attempt is made to eliminate<br />

these toxic materials before final disposal occurs. Some scientists have worried<br />

that food crops grown in soil fertilized by sewage sludge may incorporate<br />

127


648 Chapter 14 The Pollution and Purification of Water<br />

some of the increased amounts of heavy metals. Control experiments indicate<br />

that vegetables vary greatly in the extent to which they will absorb increased<br />

amounts of the metals; e.g., the uptake of lead by lettuce is particularly large,<br />

but that by cucumbers is negligible. The concentration of arsenic in agricultural<br />

soils is greatly increased if arsenic pesticides are applied to them; crops<br />

planted on these soils subsequently absorb some of the adsorbed arsenic.<br />

Other substances of concern in using sewage sludge as fertilizer for food are<br />

alkylphenols from detergents, brominated fire retardants, and pharmaceuticals—<br />

especially antibiotics given to farm animals.<br />

Modern Wastewater and<br />

Air Purification Techniques<br />

The most important chemical (as opposed to biological) pollutants dissolved<br />

in wastewater are usually chloroorganics, phenols, cyanides, and heavy metals.<br />

Below we describe some of the high-tech methods that have recently been<br />

developed and put into practice to purify wastewater, particularly for removal<br />

of chloroorganics. Some of these same techniques are also used to cleanse the<br />

compounds from contaminated air.<br />

The Destruction of Volatile Organic Compounds<br />

The major stationary sources in North America of VOCs (Chapter 3) are the<br />

evaporation of organic solvents, the manufacture of chemicals, and the petroleum<br />

industry and its storage activities. Wastewater effluent that is contaminated<br />

with VOCs, e.g., the water emanating from chemical or petrochemical<br />

plants, is commonly treated by a two-step process:<br />

1. The VOCs are removed from the wastewater by air stripping. In this<br />

process, air is passed upward into a downward stream of the water, and<br />

the volatile materials are transferred from the liquid to the gas phase. This<br />

technique does not work well for compounds that are highly water soluble.<br />

2. The resulting VOCs, now present in low concentration in a contained<br />

mass of humid air, are destroyed by a process of catalytic oxidation. For<br />

example, air heated to 300–500°C is passed for a short time over platinum<br />

or, depending upon the VOC, some other precious metal that is supported<br />

on alumina. The energy costs of this step are very high since it involves<br />

heating a large volume of humid air. Note that the outlet air from such<br />

processes contains hydrogen chloride, HCl, if the VOCs originally contained<br />

chlorine; this compound must be removed by scrubbing with a basic substance<br />

before the air is released into the atmosphere.<br />

The removal of VOCs from gaseous emissions from industries usually<br />

operates by the same catalytic oxidation process; typically the concentration<br />

128


Modern Wastewater and Air Purification Techniques 649<br />

of VOCs in the air stream is thereby reduced by 95%. A primary heat<br />

exchanger recovers and reuses the VOCs’ heat of combustion to warm<br />

incoming gases to the operating temperature.<br />

The adsorption of compounds onto activated carbon (see Box 14-1) or<br />

onto synthetic carbonaceous adsorbents is a cost-effective technology used<br />

for the removal of low-level VOC concentrations from both liquid and vapor<br />

streams; it is also useful for nonvolatile organic compounds. These adsorbents<br />

can be easily regenerated by treatment with steam or by other thermal techniques<br />

as well as by solvents; the concentrated pollutants can be subsequently<br />

destroyed by catalytic oxidation.<br />

Advanced Oxidation Methods for Water Purification<br />

Conventional water purification methods often do not successfully deal with<br />

synthetic organic compounds such as chloroorganics that are dissolved at low<br />

concentrations; examples include the common groundwater pollutants<br />

trichloroethene and perchloroethene. The conventional method for the<br />

treatment of water containing such pollutants is adsorption of the chloroorganics<br />

onto activated carbon; this removes the compounds but doesn’t<br />

destroy them. The wastewater from pulp-and-paper mills also contains<br />

organochlorines that are resistant to conventional treatments.<br />

In order to cleanse water of these extra-stable organics, so-called<br />

advanced oxidation methods (AOMs) have been developed and deployed.<br />

The aim of these methods is to mineralize the pollutants, i.e., to convert<br />

them entirely to CO2 , H2O, and mineral acids such as HCl. Most AOMs are<br />

ambient-temperature processes that use energy to produce highly reactive<br />

intermediates of high oxidizing or reducing potential, which then attack and<br />

destroy the target compounds. The majority of the AOMs involve the<br />

generation of significant amounts of the hydroxyl free radical, OH, which<br />

in aqueous solution is a very effective oxidizing agent, as it is in air (see<br />

Chapters 1–5). The hydroxyl radical can initiate the oxidation of a molecule<br />

by extraction of a hydrogen atom or addition to one atom of a multiple bond,<br />

as it does in air (Chapter 5); in water, as an additional alternative, it can also<br />

extract an electron from an anion.<br />

Since the generation of OH in solution is a relatively expensive process,<br />

it is economical to use AOMs to treat only the components of the wastes that<br />

are resistant to the cheaper, conventional treatment processes. Thus, integrating<br />

an AOM with pretreatment of the wastewater by biological or other<br />

processes to first dispose of the easily oxidized materials is often appropriate.<br />

Ultraviolet (UV) light is often used to initiate the production of hydroxyl<br />

radicals and thus to begin the oxidations. Commonly, hydrogen peroxide,<br />

H2O2, is added to the polluted water and UV light from a strong source in the<br />

200–300-nm range is shone on the solution. The hydrogen peroxide absorbs<br />

the ultraviolet light (especially that closer to 200 nm than to 300 nm) and<br />

129


650 Chapter 14 The Pollution and Purification of Water<br />

uses the energy obtained to split the O9O bond, resulting in the formation<br />

of two OH radicals:<br />

UV<br />

H 2O 2 9: 2 OH<br />

Alternatively, and less commonly, ozone is produced and then photochemically<br />

decomposed by UV light. The resulting oxygen atom reacts with water<br />

to efficiently produce OH via the intermediate production of hydrogen peroxide,<br />

which is photolyzed:<br />

UV<br />

O 3 9: O 2 * O *<br />

UV<br />

O * H 2 O 9: H 2 O 2 9: 2 OH<br />

A fraction of the oxygen atoms produced by ozone photolysis are electronically<br />

excited, and these react with water to directly produce hydroxyl radicals,<br />

as discussed in Chapter 1.<br />

PROBLEM 14-15<br />

Given that the enthalpies of formation for H 2 O 2 and OH are, respectively,<br />

136.3 and 39.0 kJ mol 1 , calculate the heat energy required to dissociate<br />

one mole of hydrogen peroxide into hydroxyl free radicals. What is the maximum<br />

wavelength of light that could bring about this transformation? [Hint:<br />

See Chapter 1]. Given that light of 254-nm wavelength is usually used, and<br />

that all the energy of each photon that is in excess of that required to dissociate<br />

one molecule is lost as waste heat, calculate the maximum percentage of<br />

the input light energy that can be used for dissociation itself.<br />

Hydroxyl radicals for wastewater treatment can also be efficiently produced<br />

without the use of UV light by combining hydrogen peroxide with<br />

ozone. The chemistry of the intermediate processes is complex, but the overall<br />

reaction between these two species is<br />

H2O2 2 O3 9: 2 OH 3 O2 This ozone/H2O2 method is more cost-effective and easier to adapt to existing<br />

water treatment systems than is any other AOM system.<br />

It is also possible to generate the hydroxyl radical electrolytically. In most<br />

such applications, a metal ion (such as Ag or Ce 3 ) is first oxidized to a<br />

more positively charged ion (Ag 2 or Ce 4 in our examples) that will subsequently<br />

oxidize water to H and OH.<br />

The biggest liability associated with advanced oxidation processes is that<br />

their action produces toxic chemical by-products. For example, in the ozone/<br />

peroxide and peroxide/UV treatments of groundwater contaminated with<br />

130


trichloroethene and perchloroethene, the toxic intermediates trichloroacetic<br />

acid, CCl 3 COOH, and dichloroacetic acid, CHCl 2 COOH, are formed in about<br />

1% yield.<br />

Photocatalytic Processes<br />

Modern Wastewater and Air Purification Techniques 651<br />

Another innovative technology for wastewater treatment involves the irradiation<br />

by UV light of solid semiconductor photocatalysts such as titanium<br />

dioxide, TiO 2, small particles of which are suspended in solution. Titanium<br />

dioxide is chosen as the semiconductor for such applications since it is nontoxic,<br />

is very resistant to photocorrosion, is cheap and plentiful, absorbs light<br />

efficiently in the UV-A region, and can be used at room temperature. Irradiation<br />

at wavelengths less than 385 nm produces electrons, e , in the conduction<br />

band and holes, h , in the valence band of the metal oxide. The holes in the<br />

semiconductor can react with surface-bound hydroxide ions or with water<br />

molecules, thereby producing hydroxyl radicals in both cases:<br />

h OH 9: OH<br />

h H 2O 9: OH H <br />

The holes can also react directly with adsorbed pollutants, producing radical<br />

cations that readily engage in subsequent degradation reactions.<br />

Normally, O 2 molecules dissolved in the water react with the electron<br />

produced at the semiconductor surface, a process that eventually produces<br />

more reactive free radicals but is relatively slow. If hydrogen peroxide is added<br />

to the water instead, it will react with the electron to form the anion radical<br />

and generate reactive radicals more quickly.<br />

The cost of the electrical energy required to generate the needed UV<br />

light is usually the major expense in operation of AOM systems. On this<br />

basis, the titanium dioxide methods are even less cost-effective than those<br />

described previously, since considerably more electricity is required per pollutant<br />

molecule destroyed. Sunlight could be used to supply the UV light, but<br />

only about 3% of its light lies in the appropriate UV-A range and is absorbed<br />

by the solid. Another problem with the TiO 2 processes is the difficulty in separating<br />

the various reactants and products from the TiO 2 particle if the metal<br />

oxide has been used in the form of a fine powder. However, there are now<br />

closed systems in which the titanium dioxide slurry is efficiently separated<br />

from the purified water and recycled back to the inlet stream.<br />

Some scientists have experimented with immobilizing TiO 2 as a thin film<br />

(1 mm thick) on a solid surface such as glass, tile, or alumina. Indeed,<br />

TiO 2 -coated tiles are now used on walls and floors in some buildings. The<br />

low-level UV light from fluorescent lighting in such rooms is sufficient to<br />

allow the destruction of gaseous and liquid-phase pollutants that touch the<br />

oxide on the tiles! For example, odors that upset people are usually present in<br />

131


652 Chapter 14 The Pollution and Purification of Water<br />

Review Questions<br />

air at concentrations of only about 10 ppm; at such levels, the UV from<br />

normal fluorescent lighting should be sufficient to destroy them with TiO 2<br />

photocatalysts. Bacterial infections such as those that cause many secondary<br />

infections in hospitals can also be eliminated by spraying walls and floors (in<br />

rooms lit by fluorescent bulbs) to give them a titanium dioxide film. Photocatalysts<br />

are quiet, unobtrusive cleansing materials.<br />

Other Advanced Oxidation Methods<br />

A process called direct chemical oxidation has been proposed for the destruction<br />

of solid and liquid organic wastes in the aqueous phase, particularly in<br />

environments such as those under buildings, where the light required for<br />

UV processes cannot conveniently be supplied. It uses one or another of<br />

the strongest known chemical oxidants—e.g., acidified peroxydisulfate<br />

2<br />

anion, S2O8 , under ambient pressure and moderate temperatures to oxidize<br />

the wastes. Such a process needs no catalysts and produces no secondary<br />

wastes of concern. The sulfate that results from the peroxydisulfate can be<br />

recycled back to the oxidant. Other very strong oxidizing agents that have<br />

<br />

been tested are the peroxymonosulfate anion, HSO5 , and the ferrate ion,<br />

2<br />

FeO4 ; in the latter, iron has an oxidation number of 6, so it is not surprising<br />

that it is a strong oxidizing agent. Unfortunately, ferrate ion suffers from<br />

the problem of instability.<br />

PROBLEM 14-16<br />

1. Describe the function of (a) aeration and<br />

(b) addition of aluminum or iron sulfate in the<br />

purification of drinking water.<br />

2. Describe the chemistry underlying the removal<br />

of excess calcium and magnesium ions from drinking<br />

water.<br />

3. Describe how water can be disinfected by<br />

(a) membrane filtration and (b) ultraviolet<br />

irradiation.<br />

Deduce the balanced half-reaction (acidic media) in which the peroxydisulfate<br />

ion is converted into sulfate ion. Repeat the exercise for the conversion of<br />

oxalic acid, C 2 H 2 O 4 , into carbon dioxide. Combine these half-reactions into<br />

a balanced equation, and calculate the volume of 0.010 M peroxydisufate that<br />

is required to oxidize one kilogram of oxalic acid.<br />

4. What two other chemical methods, other than<br />

chlorination, are used to disinfect water? What are<br />

some advantages and disadvantages to these<br />

alternatives?<br />

5. Explain the chemistry underlying the disinfection<br />

of water by chlorination. What is the<br />

active agent in the destruction of the pathogens?<br />

What are the practical sources of the active<br />

ingredient?<br />

132


6. Explain why pH control of water in swimming<br />

pools is important. What compounds are formed<br />

when the chlorinated water reacts with ammonia?<br />

7. Discuss the advantages and disadvantages of<br />

using chlorination to disinfect water, including the<br />

nature of the THM compounds.<br />

8. What is meant by the terms groundwater and<br />

aquifer? How does the saturated zone of soil differ<br />

from the unsaturated?<br />

9. Why did concern about groundwater pollution<br />

lag far behind that about surface water?<br />

10. Name three important sources of nitrate ion<br />

to groundwater.<br />

11. Construct a table that shows the common<br />

oxidation numbers for nitrogen. Deduce in which<br />

column the following environmentally important<br />

compounds belong: HNO2 , NO, NH3 , N2O, N2 ,<br />

<br />

HNO3, NO3 . Which of the species become<br />

prevalent in aerobic conditions in a lake? Under<br />

anaerobic conditions? What is the oxidation<br />

number of nitrogen in NH2OH? 12. Explain why excess nitrate in drinking water<br />

or food products can be a health hazard; include<br />

the relevant balanced chemical reaction showing<br />

how nitrate becomes reduced.<br />

13. What is an N-nitrosamine? Write the structure<br />

and the full name for NDMA.<br />

14. What is the formula for the perchlorate ion?<br />

What is the origin of perchlorate ion in U.S.<br />

drinking water?<br />

15. Define leachate.<br />

16. Name two types of organic contaminants<br />

found in groundwater, and give two examples of<br />

each type.<br />

17. Explain the difference in vertical location in<br />

an aquifer between compounds such as chloroform<br />

and those such as toluene.<br />

133<br />

Review Questions 653<br />

18. Define the term plume and describe how it<br />

forms in an aquifer.<br />

19. Why are the BTX and MTBE components<br />

of gasoline the ones that are most often found<br />

in groundwater? Are both components easily<br />

biodegraded?<br />

20. What is meant by reductive degradation?<br />

Describe the in situ technique by which chloroorganics<br />

in aquifers can be destroyed by reductive<br />

dechlorination.<br />

21. What procedures are involved in primary<br />

wastewater treatment? In secondary treatment?<br />

22. List five possible water purification processes<br />

that are associated with the tertiary treatment<br />

of wastewater, including one that removes<br />

phosphate ion.<br />

23. What polyphosphate was commonly used in<br />

detergents, and why did its use lead to environmental<br />

problems? What are the other main sources<br />

of phosphate to natural waters? What other<br />

builders are used in detergents?<br />

24. Describe two important methods that are used<br />

to desalinate wastewater.<br />

25. Describe the chemical processes by which<br />

cyanide ion can be removed from wastewater.<br />

26. Describe how VOCs dissolved in wastewater<br />

are usually removed and destroyed.<br />

27. Describe what AOMs stands for, and state the<br />

most common reactive agent in such processes.<br />

Describe three methods by which this reactive<br />

species can be generated.<br />

28. Describe two photocatalytic methods that can<br />

destroy organic wastes.<br />

29. What is direct chemical oxidation? What are<br />

two of the strong oxidizing agents that can be used<br />

for such procedures?


654 Chapter 14 The Pollution and Purification of Water<br />

Green Chemistry Questions<br />

See the discussion of focus areas and the principles<br />

of green chemistry in the Introduction before<br />

attempting these questions.<br />

1. What are the environmental advantages of using<br />

iminodisuccinate compared to most chelating agents?<br />

2. The development of iminodisuccinate by<br />

Bayer won a Presidential Green Chemistry<br />

Challenge Award.<br />

Additional Problems<br />

1. Given that for HOCl, K a 2.7 10 8 , deduce<br />

the fraction of a sample of the acid in water that<br />

exists in the molecular form at pH values (predetermined<br />

by the presence of other species) of 7.0, 7.5,<br />

8.0, and 8.5. [Hint: Derive an expression that relates<br />

the fraction of HOCl that is ionized to the concentration<br />

of hydrogen ions.] Would it be a good idea to allow<br />

the pool water’s pH to rise to 8.5?<br />

2. The equilibrium constant for the reaction of<br />

dissolved molecular chlorine with water to give<br />

hydrogen ions, chloride ions, and HOCl is 4.5 <br />

10 4 , where as usual the concentration of water is<br />

included in the K value. If the pH of the solution is<br />

determined by other processes so that the amount<br />

of hydrogen ion contributed by the chlorine reaction<br />

is negligible, calculate the fraction of the original<br />

50 ppm chlorine which remains as Cl 2 at pH<br />

values of 0, 1, and 2. [Notes: (1) The dissociation of<br />

HOCl into ions is negligible at these low pH values.<br />

(2) Approximate solutions to the quadratic<br />

equation involved in these calculations will not be<br />

accurate due to the high percentage of reaction.]<br />

3. Calculate the volume of Ca 5(PO 4) 3OH, the<br />

density of which is 3.1 g/mL, which is produced for<br />

each gram of sodium tripolyphosphate present in a<br />

detergent when it is removed in tertiary wastewater<br />

treatment. Estimate the annual mass of detergent<br />

used for laundry purposes for a typical household of<br />

four persons. Assuming that the phosphate levels in<br />

(a) Into which of the three focus areas for these<br />

awards does this award best fit?<br />

(b) List at least three of the twelve principles of<br />

green chemistry that are addressed by the green<br />

chemistry developed by Bayer.<br />

laundry detergents used were about 50%, calculate<br />

the volume that was required annually to dispose<br />

of its waste laundry phosphate.<br />

4. Calculate the oxidation number of the<br />

chlorine in molecular chlorine, HOCl, chlorine<br />

dioxide, monochloramine (NH 2Cl), and<br />

sodium chloride. Given that the last item is<br />

the most stable form of chlorine, predict whether<br />

the other substances mentioned are likely to be<br />

oxidizing agents or reducing agents, and rank<br />

them in likely order of this redox behavior. Using<br />

this analysis and the section on chlorine compounds<br />

in your introductory chemistry textbook,<br />

suggest other compounds that might be useful to<br />

disinfect water.<br />

5. Given their names, can you deduce the nature<br />

of the similarity in molecular structure between<br />

hydrogen peroxide and the sulfur compounds that<br />

are used in direct chemical oxidation methods?<br />

By calculating the oxidation number of the atoms<br />

in hydrogen peroxide, deduce why it can act as an<br />

oxidizing agent.<br />

6. What could be done to dispose of the solvents<br />

that are used to extract VOCs from adsorbents?<br />

7. Write the initial reaction step that occurs if<br />

methyl chloroform, CH 3CCl 3, were to be<br />

destroyed by (a) reductive degradation and (b)<br />

hydroxyl radical attack.<br />

134


8. Water samples from three wastewater streams<br />

were analyzed, and the important pollutants<br />

determined to be those listed below. In each case,<br />

devise economical, practical processes (other than<br />

activated carbon treatment) for purifying the water<br />

of the three pollutants:<br />

(a) Phosphate ion, ammonium ion, and salt<br />

(in water containing bicarbonate ion)<br />

(b) Nitrite ion, PCE, and Fe(II)<br />

(c) Cadmium ion, carbon tetrachloride, and glucose<br />

9. Write the reaction that occurs between hydroxyl<br />

radical and carbonate ion dissolved in water. What<br />

are two alternative substances that you could add to<br />

the water to decrease the carbonate ion concentration,<br />

one that operates by elimination of carbon dioxide<br />

and the other by precipitation of the ion, so as to<br />

decrease the amount of hydroxyl radical destroyed by<br />

this reaction? Given the solubility product constants<br />

in this chapter, estimate the lowest practical carbonate<br />

concentration you could reach by this process,<br />

and comment on the desirability of the ions that you<br />

have introduced into the water.<br />

10. Treated drinking water should contain<br />

0.5 mg/L of Cl 2 after most of the chlorine has<br />

been converted to HOCl. What pressure of Cl 2(g)<br />

is required to maintain this concentration?<br />

K H 8.0 10 3 M for Cl 2.<br />

Additional Problems 655<br />

11. At a particular temperature, K a for HOCl is<br />

3.5 10 8 . What would be the pH values for<br />

1.00 M and 0.100 M concentrations of HOCl at<br />

this temperature? What percentages of the HOCl is<br />

undissociated at these two concentrations?<br />

12. To desalinate seawater by reverse osmosis,<br />

pressure in excess of a solution’s osmotic pressure,<br />

p, must be applied across the membrane. The total<br />

osmotic pressure exerted in a solution is determined<br />

by the total molar concentration, M, of its solutes,<br />

and is given by the equation p MRT. Using the<br />

composition of seawater listed in Chapter 13, determine<br />

the minimum pressure that must be exerted<br />

on seawater to desalinate it using reverse osmosis at<br />

20°C. Recall that R 0.082 L atm mol 1 K 1 .<br />

13. Chlorine-containing substances covering a wide<br />

variety of oxidation numbers have been encountered,<br />

some within this chapter and others previously<br />

in the book. For each substance in the list<br />

below, write out its formula, deduce the oxidation<br />

number of its chlorine, and fill in the appropriate<br />

row of the table below, as in the example shown:<br />

chlorite ion perchlorate ion<br />

molecular chlorine hypochlorous acid<br />

chlorine dioxide chlorine monoxide<br />

chloride ion chlorate ion<br />

Oxidation Number Formula of Example Name of Example<br />

1<br />

0<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

ClO3 Chlorine trioxide<br />

135


656 Chapter 14 The Pollution and Purification of Water<br />

Further Readings<br />

1. A. Kolch, “Disinfecting Drinking Water with UV<br />

Light,” Pollution Engineering (October 1999): 34.<br />

2. F. Bove et al., “Drinking Water Contaminants<br />

and Adverse Pregnancy Outcomes: A Review,”<br />

Environmental Health Perspectives 110 (supplement 1)<br />

(2002): 61.<br />

3. J. R. Nuckols et al., “Influence of Tap Water<br />

Quality and Household Use Activities on Indoor<br />

Air and Internal Dose Levels of Trihalomethanes,”<br />

Environmental Health Perspectives 113 (2005): 863.<br />

See also ibid, 114 (2006): 514.<br />

4. U. van Gunten, “Ozonation of Drinking Water,”<br />

Parts I and II, Water Research 37 (2003): 1443, 1469.<br />

5. S. D. Richardson et al., “Identification of New<br />

Ozone Disinfection Byproducts in Drinking<br />

Water,” Environmental Science and Technology 33<br />

(1999): 3368.<br />

6. R. F. Service, “Desalination Freshens Up,”<br />

Science 313 (2006): 1088.<br />

7. “Chlorinated Solvent Source Zones,” Environmental<br />

Science and Technology (June 1, 2003): 225A.<br />

8. P. B. Hatzinger, “Perchlorate Biodegradation for<br />

Water Treatment,” Environmental Science and Technology<br />

(June 1, 2005): 239A.<br />

9. L. Fewtrell, “Drinking-Water Nitrate, Methemoglobinemia,<br />

and Global Burden of Disease: A<br />

Discussion,” Environmental Health Perspectives 112<br />

(2004): 1371.<br />

10. M. P. Zeegers et al., “Nitrate Intake Does Not<br />

Influence Bladder Cancer Risk: The Netherlands<br />

Cohort Study,” Environmental Health Perspectives<br />

114 (2006): 1527.<br />

11. M. A. Montgomery and M. Elimelech,<br />

“Water and Sanitation in Developing<br />

Websites of Interest<br />

Log on to www.whfreeman.com/envchem4/ and click on Chapter 14.<br />

Countries,” Environmental Science and Technology<br />

41 (2007): 17.<br />

12. B. T. Nolan et al., “Risk of Nitrate in Groundwaters<br />

of the United States—A National Perspective,<br />

Environmental Science and Technology 31<br />

(1997): 2229.<br />

13. J. A. MacDonald, “Evaluating Natural Attenuation<br />

for Groundwater Cleanup,” Environmental<br />

Science and Technology (August 1, 2000): 346A.<br />

14. L. W. Canter, R. C. Knox, and D. M.<br />

Fairchild, Groundwater Quality Protection (Boca<br />

Raton, FL: Lewis Publishers, 1987).<br />

15. E. K. Nyer, Groundwater Treatment Technology,<br />

2nd ed. (New York: Van Nostrand Reinhold, 1992).<br />

16. D. M. Mackay and J. A. Cherry, “Groundwater<br />

Contamination: Pump-and-Treat Remediation,”<br />

Environmental Science and Technology 23<br />

(1989): 630.<br />

17. R. J. Gilliom, “Pesticides in U.S. Streams<br />

and Groundwater,” Environmental Science and<br />

Technology 41 (2007): 3409.<br />

18. “Environmental Processes ’96: A Special<br />

Report,” Hydrocarbon Processing Magazine (International<br />

Edition) 75 (1996): 85 [reviews many emerging<br />

technologies that can handle water and air<br />

pollution problems].<br />

19. D. Simonsson, “Electrochemistry for a cleaner<br />

environment,” Chemical Society Reviews 26 (1997):<br />

181.<br />

20. N. C. Baird, “Free Radical Reactions in Aqueous<br />

Solutions: Examples from Advanced Oxidation<br />

Processes for Wastewater and from the<br />

Chemistry in Airborne Water Droplets,” Journal of<br />

Chemical Education 74 (1997): 817.<br />

136


Environmental<br />

Instrumental<br />

Analysis V<br />

Ion Chromatography of Environmentally Significant Anions 657<br />

The quantitative determination of levels of environmentally<br />

important ions, such as those discussed in<br />

the preceding chapters, can be accomplished using<br />

chromatographic methods described in this box.<br />

The need to determine the prevalence of<br />

3<br />

common anions like phosphate (PO4 ),<br />

<br />

nitrate (NO3 ), or fluoride (F ) isn’t immediately<br />

clear. The biospheric significance of<br />

these ubiquitous ions is not as obvious as is, for<br />

example, the presence of PCBs, pesticides, or<br />

toxic metals like lead, mercury, or cadmium.<br />

These ionic components are important<br />

because they give an indication of the relative<br />

reduction–oxidation potential in an aqueous<br />

sample taken from an environment such as a<br />

3<br />

stagnant lake (PO4 ), or of the contamination<br />

of groundwater from fertilizer runoff<br />

(NO 3<br />

), or of whether municipal water sup-<br />

plies need to be supplemented with fluoride<br />

(F ) for the health of children’s teeth.<br />

Although these charged ions can be detected<br />

by widely available ultraviolet detectors<br />

common in most high-performance liquid<br />

chromatographic systems (Janos and Aczel,<br />

1996), a more sensitive means of detection<br />

involves ionic conductivity. This chromatographic<br />

method is called ion chromatography<br />

with ionic conductivity detection. Although<br />

cations can also be separated by ion chromatography<br />

(IC), only anionic separations<br />

will be discussed here.<br />

The heart of the separation process in<br />

an ion chromatograph is a short column<br />

(10–15 cm) packed with small-diameter particles<br />

called ion exchange resins. These are often<br />

made of a styrene/divinylbenzene polymer or<br />

Ion Chromatography of Environmentally<br />

Significant Anions<br />

137<br />

microparticles of silica coated with compounds<br />

containing an anionic functional group such<br />

<br />

as a quaternary amine, ¬N(CH3) 3 OH ,ora<br />

<br />

primary amine, ¬NH3 OH , when they are<br />

to be used for anion separation.<br />

The actual process of chromatographic<br />

separation occurs after a sample containing<br />

analyte anions (and their associated cations)<br />

is injected onto the chromatographic column.<br />

With gas chromatography (see Environmental<br />

Instrumental Analysis Box II), the mobile<br />

phase is an inert gas that does not chemically<br />

interact with the chromatographic surface.<br />

The mobile phase in ion chromatography, on<br />

the other hand, is a solution of cations and<br />

anions with a carefully controlled pH; often<br />

buffers are used. This complex mixture of<br />

mobile-phase ions—carefully chosen for each<br />

group of analyte ions to be separated—interacts<br />

with the analyte ions and the functional<br />

groups of the column’s chromatographic<br />

surface. That interaction involves competition<br />

of the mobile-phase anions and the analyte<br />

anions for chromatographic sites on the<br />

packing material (the charged functional<br />

<br />

groups such as ¬N(CH3) 3 or ¬SO3 ). This<br />

competition yields different overall travel<br />

times for each of the analytes as they pass down<br />

the column; some are retained longer than<br />

others. (The overall down-column movement<br />

is provided by a pumping of the mobile<br />

phase by an external pump.) Different analyte<br />

travel times—as in gas chromatography—<br />

translate into different exit (or retention)<br />

times for each anion in the original mixture.<br />

The result is chromatographic separation of<br />

anions.<br />

(continued on p. 658)


658 Environmental Instrumental Analysis V<br />

Environmental<br />

Instrumental<br />

Analysis V<br />

The process for anionic ion chromatographic<br />

retention by ion exchange resins can<br />

be represented by the equation<br />

<br />

RN(CH3) 3 HCO3 (s) anion (aq) 9:<br />

<br />

RN(CH3) 3 anion (s) HCO3<br />

Ion Chromatography of Environmentally<br />

Significant Anions (continued)<br />

(aq)<br />

In this equation, the term anion represents<br />

any of the analyte anions mentioned above.<br />

When a sample is injected onto the column,<br />

this anion is quickly retained by complexation<br />

with the stationary phase near the head of the<br />

column. The next step in the chromatographic<br />

process takes place as a mobile phase, with a<br />

carefully controlled amount of anionic ion<br />

<br />

such as bicarbonate, HCO3 , is pumped<br />

through the column. The presence of the bicarbonate<br />

anion in the mobile phase forces the<br />

equilibrium in the above equation to the left;<br />

the retained analyte anion is freed and moves<br />

down the column in the flowing mobile phase.<br />

As the analyte moves along, it repeatedly<br />

undergoes this same process of retention and<br />

movement (or exchange between the stationary<br />

and mobile phases). Most importantly, different<br />

analyte anions (e.g., fluoride or phosphate<br />

or chloride) undergo this exchange process to<br />

differing degrees and therefore travel at different<br />

overall rates during their time in the IC<br />

column. The result is that different analytes<br />

exit the chromatographic column at different<br />

times, i.e., separation has taken place.<br />

The task of detecting analyte anions in<br />

the presence of the anions always present in<br />

the mobile phase is by no means a trivial one.<br />

Since both kinds of anions—the analytes’ and<br />

the mobile phase’s—conduct electricity, using<br />

an ordinary conductivity cell as a detector at<br />

the end of the IC column is normally not<br />

practical. The problem is especially difficult<br />

because, in order to obtain adequate separation<br />

of some important anions, the mobile<br />

phase often has to have high ionic content to<br />

displace the analyte anions from the chromatographic<br />

surface—something that is<br />

obviously required for separation. Therefore,<br />

most of the ionic conductivity passing through<br />

the detector is attributable to the mobilephase<br />

ions and not the analyte—an unworkable<br />

situation when trying to detect the<br />

analyte anions by their conductivity.<br />

An ingenious solution to this problem is<br />

called conductivity suppression or eluent (mobilephase)<br />

suppression. This technology converts<br />

the mobile-phase anions from an easily dissociated<br />

ionic form to a (soluble) molecular form<br />

that does not strongly influence the signal<br />

produced by the conductivity detector. The<br />

suppression module is placed after the chromatographic<br />

column but before the conductivity<br />

detector. In an anion exchange system,<br />

the suppression module might carry out the<br />

following reaction:<br />

Na <br />

(aq) HCO3 (aq) resin H (s) 9:<br />

resin – Na (s) H2CO3 (aq)<br />

Here resin H represents a cation exchange<br />

resin that will exchange cations—instead of<br />

anions as in the chromatographic column<br />

described. This process basically prevents (or<br />

suppresses) the mobile phase’s anions from<br />

contributing to the conductivity by converting<br />

current-conducting bicarbonate anion into<br />

relatively undissociated H 2CO 3. Therefore, the<br />

conductivity detector’s signal is based almost<br />

138


Mobilephase<br />

pump<br />

Ion Chromatography of Environmentally Significant Anions 659<br />

Injector<br />

Sample is<br />

introduced.<br />

Ion exchange<br />

column<br />

Analytes’<br />

retention times<br />

based on<br />

interaction<br />

of analyte,<br />

ion exchange<br />

column, and<br />

mobile-phase<br />

ions.<br />

completely on the passage of analyte anions<br />

through the detector cell. (Those anions are<br />

not affected by the cation ion exchange resin.)<br />

This results in lower (i.e., better) detection<br />

limits for the analytes of interest and a more<br />

stable baseline (less noise and drift) than a<br />

similar system without eluent suppression.<br />

The figure above is a schematic of an ion<br />

chromatographic system. Detailed are an injector,<br />

chromatographic column, eluent suppression<br />

module, and conductivity detector, as well<br />

as the processes that occur at each step.<br />

The figure at right is an example of the<br />

kind of chromatogram that a system of this<br />

type would generate. The anions detected are<br />

fluoride, chloride, phosphate, and nitrate. As<br />

with all chromatograms, detector signal intensity<br />

is plotted versus time.<br />

Researchers have used this method<br />

recently to determine nitrate and nitrite anions<br />

in dew, rain, and snow collected in Massachusetts<br />

(Zuo et al., 2006). Instead of conductivity<br />

detection, these authors used a UV absorption<br />

139<br />

Eluent<br />

suppression<br />

module<br />

Mobile-phase<br />

ions are<br />

converted<br />

to molecular<br />

forms that do<br />

not produce<br />

significant<br />

signals<br />

in detector.<br />

Conductivity<br />

detector<br />

Analyte ions<br />

are detected<br />

against a quiet,<br />

stable background<br />

by<br />

their electrical<br />

conductivity.<br />

detector and chose an analytical wavelength at<br />

which neither the mobile phase nor other<br />

anions would absorb (205 nm). Surprisingly,<br />

dew had the highest concentrations of these<br />

Detector signal<br />

Fluoride<br />

Chloride<br />

Phosphate<br />

Nitrate<br />

0 4 8 12 16 20<br />

Time (minutes)<br />

(continued on p. 660)


660 Environmental Instrumental Analysis V<br />

Environmental<br />

Instrumental<br />

Analysis V<br />

* Detection limit for nitrite 10 ppb.<br />

anions compared to rain and snow collected at<br />

the same site (see table above). The authors proposed<br />

that these dew nitrate concentrations,<br />

ranging from 4.79 to 5.99 g/mL, suggest that<br />

dew is acting as a nighttime sink for these<br />

anionic species; they also note that this may be<br />

important for vegetation because these anions<br />

are held in contact with the leaf surface for long<br />

time periods as dew forms, and the concentration<br />

may spike as dew evaporates in the morning.<br />

Since photolysis of both these anions in<br />

shallow aqueous solution can lead to the formation<br />

of hydroxyl radical and hydrogen peroxide,<br />

this may be a source of oxidative stress<br />

for plants on which the dew forms (Kobayashi<br />

et al., 2002):<br />

<br />

NO2 H2O light 9: OH NO OH <br />

NO 3 H2O light 9: OH NO 2 OH <br />

Studies of this process on red pine needles on<br />

trees on Mount Gokurakuji in western Japan<br />

Ion Chromatography of Environmentally<br />

Significant Anions (continued)<br />

Sample Date Nitrite (ppb) Nitrate (ppm)<br />

Dew 1 27.09.2005 640 4.87<br />

2 27.09.2005 620 4.79<br />

3 26.09.2005 830 5.99<br />

Rain 1 02.10.2005 < DL* 2.63<br />

2 02.09.2005 < DL* 2.62<br />

3 28.05.2005 140 1.20<br />

Snow 1 29.01.2005 21 0.320<br />

2 18.01.2005 32 0.376<br />

3 12.01.2005 32 0.60<br />

4 12.01.2005 26 0.56<br />

have concluded that 40% of hydroxyl radial<br />

production in dew on those trees originates<br />

from nitrite and nitrate (Nakatani et al.,<br />

2001).<br />

References: P. Janos and P. Aczel, “Ion Chromatographic<br />

Separation of Selenate and Selenite Using a<br />

Polyanionic Eluent,” Journal of Chromatography A 749<br />

(1996): 115–122.<br />

T. Kobayashi, N. Natanani, T. Hirakawa, M. Suzuki,<br />

T. Miyake, M. Chiwa, T. Yuhara, N. Hashimoto,<br />

K. Inoue, K. Yamamura, N. Agus, J. R. Sinogaya,<br />

K. Nakane, A. Kume, T. Arakaki, and H. Sakugawa,<br />

Environmental Pollution 118 (2002): 383–391.<br />

N. Nakatani, T. Miyake, M. Chiwa, M. Hashimoto,<br />

T. Arakaki, and H. Sakugawa, “Photochemical Formation<br />

of OH Radicals in Dew Formed on the Pine<br />

Needles at Mt. Gokurakuji,” Water, Air and Soil<br />

Pollution 130 (2001) 397–402.<br />

Y. Zuo, C. Wang, and T. Van, “Simultaneous Determination<br />

of Nitrite and Nitrate in Dew, Rain, Snow and<br />

Lake Water Sample by Ion-Pair High-Performance<br />

Liquid Chromatography,” Talanta (2006): 281–285.<br />

140


BYLAAG 4<br />

Matthews, P, 1992, Advanced chemistry, Cambridge: Chennaice: Oxford, p. 291 – 303, 518 – 524.<br />

141


142


143


144


145


146


147


148


149


150


151


152


153


154


155


156


157


158


159


160


161


BYLAAG 5<br />

Hill, JW, Petrucci, RH, McCreary, TW and Perry, SS, 2005, General chemistry, 4 th ed., Pearson: Upper<br />

Saddle River, NY, p. 996 – 1003.<br />

162


163


164


165


166


167


168


169


170

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!