Documente online.
Zona de administrare documente. Fisierele tale
Am uitat parola x Creaza cont nou
 HomeExploreaza
upload
Upload




NANOCOMPOZITE POLIMERICE

Chimie




NANOCOMPOZITE POLIMERICE

Cuprins

ABREVIERI

REZUMAT

1. OBTINEREA NANOCOMPOZITELOR POLIMERICE

1.1. Introducere

1.2. Clasificarea nanocompozitelor polimerice

1.3. Tipuri de polimeri utilizati pentru obtinerea nanocompozitelor polimerice

1.4. Nanocompozite pe baza de polimeri biodegradabili utilizabili in aplicatii biomedicale

1.5. Metode de obtinere a nanocompozitelor polimerice pe baza de nanoargile

1.5.1. Metode Intercalative

A-Inserarea lanturilor macromoleculare sau a unui prepolimer prin intercalarea din solutie in silicati stratificati

B- Intrcalarea din topitura

1.5.2. Metode "In Situ" (cresterea lanturilor macromoleculare prin polimerizare in situ in silicati stratificati)

2. CORELAREA STRUCTURA-PROPRIETATI

2.1. Caracterizarea fizico-chimica a materialelor

Metodele (fizice si sau chimice) de analiza

2.2. Studiu comparativ polimeri-nanocompozite

2.2.1. Influenta nanoumplurii asupra proprietatilor fizico-chimice

2.2.2. Biodegradabilitatea si nanocompozite biodegradabile utilizabile in aplicatii biomedicale

ABREVIERI

ABS - acrilonitrilbutadienstiren

HDPE - polietilena de densitate mare (inalta)

Li+ - ionul de litiu

MMA - monomer metilmetaacrilat

MMT - montmorillonit

OMLS - silicati stratificati modificati organic

PA - poli acrilat

PC - policarbonat

PE - polietilena

PES - polietersulfona

PETF - polietilentereftalat

PHB - polihidroxibutirat

PHV - polihidroxivalerat

PLA - polilactida

PLS - polimer/silicat stratificat

PP - polipropilena

PMMA - polimetilmetaacrilat

PS - polistiren

PVC - policlorura de vinil

SAN - copolimer stiren-acrilonitril

REZUMAT

Referatul prezinta in prima parte stadiul actual in domeniul obtinerii si caracterizarii materialelor nanocompozite polimerice. Deasemenea sunt aratate principalele aplicatii ale acestor tipuri de materiale.

In partea a doua a referatului sunt prezentate corelatiile dintre structura si proprietatile materialelor nanocompozite polimerice. Deasemenea in aceasta parte sunt prezentate influenta nanoumpluturii (incluziunilor nanometrice) asupra proprietatilor fizico-chimice si biodegradabilitatea nanocompozite utilizabile in aplicatii biomedicale.

Cuvinte cheie: polimer, nanocompozit, incluziune, biodegradabil, aplicatii biomedicale, transport medicamente, implanturi.

1. OBTINEREA NANOCOMPOZITELOR POLIMERICE

1.1. Introducere

In ultimii ani au atras un interes deosebit din partea comunitatilor stiintifica si industriala obtinerea si caracterizarea structurilor nanocompozite polimerice, precum si descoperirea unor noi aplicatii ale acestora. Acest interes deosebit se datoreaza proprietatilor remarcabile ale materialelor nanocompozite polimerice daca le comparam cu polimerii actuali si cu macro- sau micro-compozitele conventionale [1]. Proprietatile nanocompozitelor polimerice (mult imbunatatite fata de materialele conventionale) se refera la elasticitatea , rezistenta mecanica , rezistenta termica la incalzire , permeabilitate scazuta pentru gaze [9-13], usurinta de a arde cu flacara (flamabilitate) [14-18] si la cresterea biodegradabilitatii

Pe de alta parte a fost manifestat un interes deosebit pentru fundamentarea teoretica si aplicatiile practice privind metodele de preparare si proprietatile acestor materiale [20-35], ele reprezentand sisteme-model unice pentru studierea structurii si dinamicii polimerilor in medii restranse sau limitate [36-42].

O metoda utilizata deseori pentru imbunatatirea proprietatilor mecanice ale materialelor polimerice nanocompozite este aceea a intari structura acestora prin includerea unor fibre, plachete (nanoplaci, nanopelete) sau particule [43]. O practica obisnuita pentru obtinerea unor proprietati deosebite este aceea de a introduce, adauga, spori numarul de fibre, plachete sau particule in matricea-suport, proces prin care se fabrica materialele compozite (cu caracteristici superioare fazelor luate individual). Prin utilizarea acestei metode se imbunatatesc proprietatile polimerilor, fara a se altera greutatea specifica (densitatea) sau caracteristica de a fi ductile . Proprietatile polimerilor se vor imbunatati chiar si in cazul unui continut scazut de filer (material de umplutura) [47,50,51].

In ultimii ani au aparut tehnici de procesare care permit obtinerea incluziunilor de dimensiuni nanometrice [43]. Incluziunile nanometrice se definesc ca fiind acele incluziuni care prezinta cel putin o dimensiune in domeniul 1-100 nm

In ultimii ani cercetatorii in domeniu au incercat diverse tehnici de obtinere a matricei polimerice nanocompozite. Printre aceste tehnici amintim amestecarea in topitura si polimerizarea "in situ". Este dificila realizarea unei tehnici universale pentru obtinerea nanocompozitelor polimerice datorita diferentelor fizice si chimice intre sisteme, precum si datorita diverselor tipuri de echipamente disponibile cercetatorilor. Astfel, aceste tehnici diferite vor determina obtinerea unor rezultate diferite

In ultima decada a secolului XX, majoritatea aplicatiilor polimerilor s-au limitat la obtinerea de ambalaje de tip plastic. Aceste matariale plastice sunt, in general, poliolefine (PP, PE, PS sau PVC). obtinute din prelucrarea chimica a combustibililor fosili. Cand materialele plastice ajung in mediul inconjurator ele reprezinta reziduri nedegradabile, constituindu-se intr-o problema de mediu la nivel global. Una din metodele des utilizate in ultimul timp pentru a reduce cantitatea de reziduri de mase plastice a fost incinerarea acestor polimeri dar produsul final este dioxidul de carbon (responsabil de efectul de sera) sau diverse alte gaze cu potential poluant. Alta metoda ar fi reciclarea acestor materiale, dar aceasta abordare este consumatoare de timp si energie (indepartarea altor reziduri, separarea pe categorii de mase plastice, spalarea, uscarea, reprocesarea etc.) iar produsul final are calitatea inferioara produsului initial. Tinand cont de aceste considerente, este necesara obtinerea unor polimeri ecologici (polimeri verzi) care sa nu contina ingredienti toxici si care sa poata fi degradati in mediul inconjurator. Din aceste motive, comunitatea stiitifica si cea industriala acorda o atentie deosebita dezvoltarii materialelor biodegradabile cu propritati controlate.

Polimerii biodegradabili se definesc ca fiind polimerii care sufera o scindare a lantului polimeric, scindare indusa microbiologic, avand ca rezultat final mineralizarea materialului respectiv [61]. Astfel, acesti polimeri isi gasesc aplicatii in dezvoltarea de materiale ecologice, cu potential poluant redus. Procesul de biodegradare este influentat de anumiti factori, cum ar fi: pH, temperatura, umiditate, grad de oxigenare sau prezenta anumitor metale.

Domeniul transportului substantelor farmacologic-active (medicamentelor) in organism se dezvolta rapid, captand tot mai mult atentia oamenilor de stiinta, a responsabililor cu promovarea farmaceutica si a patronatelor industriale [62]. Visul actual al cercetatorilor farmacisti este de a realiza sisteme de transport eficiente si precise al medicamentelor catre locurile de actiune ale acestora in organism [62]. Una din metodele de interes pentru transportul medicamentelor in organism este metoda incapsularii substantei farmacologic-actice in nanoparticule cu dimensiuni sub 100 nm [62]. Chiar daca aceasta nanotehnologie este la inceput, deja s-au proiectat sisteme nanocompozite de transport bazate pe nanocapsule, sisteme micelare sau nanoparticule [63,64]. Unul dintre marile avantaje ale acestor sisteme sub-micrometrice este acela ca determina o penetrabilitate intracelulara superioara particulelor micrometrice

1.2. Clasificarea nanocompozitelor polimerice

Nanotehnologia poate fi definita ca: "realizarea, procesarea, caracterizarea si utilizarea materialelor, dispozitivelor si sistemelor cu dimensiuni cuprinse in domeniul 0,1-100 nm care prezinta proprietati fizice, chimice si biologice noi, superioare datorita dimensiunilor nanometrice" [65,66]. Interesul actual in nanotehnologie este reprezentat de nano-biotehnologie, nano-sisteme, nano-electronica si materiale nano-structurate (in care se includ, ca parte semnificativa, nano-compozitele)

Materialele utilizate ca adaos, aditiv, intaritor sau umplere (reinforcement) la producerea nanocompozitelor pot fi particule (metale, combinatii ale siliciului si alti compusi anorganici sau organici), materiale stratificate (grafit, silicati stratificati sau alte minerale stratificate) sau materiale fibroase (nanofibre sau nanotuburi) (fig. 1).

Fig. 1 Cle trei tipuri principale de nano-materiale folosite la obtinerea nanocompozitelor

In functie de materialele utilizate ca adaos (aditiv de umplutura) pentru intarire (termen original in engleza - reinforcement), nanocompozitele se impart in:

Nanocompozite bazate pe nanoparticule

Nanocompozite bazate pe nanoplachete (nanostratificate)

Nanocompozite bazate pe nanofibre.

1.2.1. Nanocompozite bazate pe nanoparticule

Compozitele bazate pe particule sunt probabil cele mai folosite in materiale din zilele noastre. De obicei, particulele sunt adaugate pentru a spori elasticitatea matricei si pentru a cresterea rezistentei. Prin reducerea dimensiunii particulelo0r catre domeniul nanometric, se pot obtine materiale noi, cu proprietati superioare celor originale.

1.2.2. Nanocompozite bazate pe nanoplachete (nanostratificate)

Cele mai utilizate plachete sunt grafitul si argila. Ca materiale brute, argila si grafitul au o structura stratificata. Pentru ca folosirea acestor materiale sa fie eficienta, straturile trebuie separate si dispersate in faza matricei (fig. 2).

Fig. 2 Morfologia nanocompozitelor polimer/argila: (a) miscibil conventional, (b) intercalat si exfoliat partial, (c) intercalat si dispersat complet (d) exfoliat si dispersat complet

1.2.3. Nanocompozite bazate pe nanofibre

Nanofibrele de carbon (crescute din faza de vapori) au fost utilizate la intarirea diferitelor tipuri de polimeri (PP, PC, nylon, PES, ABS, PETF etc.). Nan 343b17d ofibrele carbonice prezinta morfologii diferite (fig. 3), de la structuri dezordonate tip lemn de bambus la structuri ordonate tip grafit cu straturi in forma de cescuta (sau cupa)

Fig. 3 Micrografiile TEM ale structurilor nanodimensionale pentru nanofibrele de carbon (a) Nanofibrele carbonice structura dezordonata tip lemn de bambus (b) si (c) structuri ordonate tip grafit cu straturi in forma de cescuta

In functie de taria interactiilor interfaciale dintre matricea polimerica si silicatul stratificat (modificat sau nu) nanocompozitele PLS se clasifica in trei categorii (fig 4), din punct de vedere termodinamic:

  1. nanocompozite intercalate pentru care indiferent de raportul dintre polimer si argila, matricea polimerica se insera in structura stratificata silicata intr+un mod regulat din punct de vedere cristalografic.
  2. nanocompozite floculate; conceptual reprezinta structuri identice cu cele intercalate, cu toate acestea straturile de silicat sunt cateodata floculate datorita interactiilor capetelor hidroxilate ale straturilor de silicat
  3. nanocompozite exfoliate pentru care straturile individuale de argila sunt separate in interiorul matricii polimerice,intre straturi existand o distanta care depinde de gradul de incarcare cu argila.

 

Nanoplacheta (nanoplacuta, foita) de argila

Intercalat Intercalat si floculat Exfoliat

 


Fig. 4 Tipurile de nanocompozite PLS

1.3. Tipuri de polimeri utilizati pentru obtinerea nanocompozitelor polimerice

Vollenberg si Heikens [52] au reusit sa produca nanocompozite de calitate buna prin amestecarea particulelor de umplutura cu matricea polimerica. Astfel, au utilizat PS, SAN, PC sau PP pentru includerea particulelor sferice de alumina de 35 sau 400 nm si includerea particulelor sferice de sticla de 4, 30 sau 100 nm. Fractia volumica a particulelor a fost cuprinsa in domeniul 0-25 %. Prepararea nanocompozitelor polimerice s-a realizat prin dizolvarea polimerului intr-un solvent polar si amestecarea nanosferelor (nanoperlelor) in acest produs timp de cateva ore.

Chan si colab. [54] au produs nanocompozite cu matrice polimerica de PP si nanoparticule de carbonat de calciu prin amestecarea in topitura a componentelor. Mai intai componentele au fost uscate intr-un cuptor la 120 °C, apoi au fost racite la temperatura camerei. Polimerul PP a fost amestecat intai cu un antioxidant. Nanoparticulele de carbonat de calciu, cu un diametru de 44 nm au fost adaugate cu grija, lent, sub agitare continua un timp stabilit. Dispersia a fost buna pentru fractii volumice ale umpluturii de 4,8 % si 9,2 %. La o fractie volumica de 13,2 % a avut loc fenomenul de agregare.

Petrovic si colab. [53] au dezvoltat un nanocompozit pe baza de poliuretan si silice (dioxid de siliciu). Silicea a fost mai intai amestecata cu poliol. Amestecul a fost apoi tratat cu diisocianat la 100 °C, timp de 16 ore in prezenta unui catalizator de Cocurre 55 aflat intr-o concentratie de 0,1 %.

Yang si colab. [56] au utilizat tehnica polimerizarii "in situ" pentru a obtine nanocompozite. Matricea este formata de poliamid-6 iar ca incluziuni s-au folosit particule de dioxid de siliciu. Dimensiunea particulelor a fost de 50 nm, obtinandu-se o buna dispersie dar prin utilizarea unor particule de 12 nm in diametru s-a produs fenomenul de agregare.

Ash si colab. si Siegel si colab. [70,71] au obtinut nanocompozite prin utilizarea polimerului PMMA ca matrice iar ca incluziuni u fost folosita pulberea de alumina. Pulberea de alumina (particulele) a fost adaugata la monomerul MMA dispersandu-se prin sonicare in solutia de vascozitate scazuta. Apoi au fost adaugate un initiator si un agent de transfer in lant. Polimerizarea a fost desfasurata sub atmosfera de azot, iar produsul final a fost fin divizat si uscat in vid.

Li si colab. [59] au utilizat o metoda originala pentru obtinerea nanocompozitelor ce sunt bazate pe HDPE si PP. S-a amestecat cele doua componente topite HDPE (75 %) si PP (25 %) apoi au fost obtinute benzi prin extrudere. Aceste benzi au fost apoi taiate in bucati mai mici si prelucrate in topitura prin extrudere sau turnare. Astfel a fost obtinut un material nanocompozit cu matrice de HDPE si nanofibre de PP cu dimensiuni de 30-150 nm in diametru.

1.4. Nanocompozite pe baza de polimeri biodegradabili utilizabili in aplicatii biomedicale

Nanotehnologia reprezinta "obtinerea, caracterizarea si utilizarea materialelor, dispozitivelor si sistemelor cu dimensiuni cuprinse in domeniul 0,1-100 nm" [65,66]. Nanotehnologia se ocupa si de obtinerea materialelor nano-structurate (in care se includ, ca parte semnificativa, nano-compozitele)

Nanocompozitele bazate pe polimeri biodegradabili si-au gasit foarte multe aplicatii datorita proprietatilor lor remarcabile.

Polimerii biodegradabilli pot fi obtinuti din surse biologice (fig. 5,6) cum ar fi: porumbul, celuloza, chitosanul, gelatina, amidonul, acidul lactic (dimerul ciclic - lactida din care se obtine PLA sau pot fi sintetizati de catre bacterii din molecule de mici dimensiuni de acid butiric sau acid valeric cand rezulta PHB sau PHV [72].

Celuloza

Chitosan

Gelatina

Fig. 5 Structurile moleculare ale celulozei si gelatinei

Alta modalitate de a obtine polimeri biodegradabili este folosirea poliesterilor alifatici si a copoliesterilor aromatici-alifatici din produsele petroliere.

Fig. 6 Structura moleculara a unui polimer tip PLA

Generatia urmatoare de materiale (materialele viitorului) este reprezentata de nanocompozitele bazate pe polimeri biodegradabili, asa-numitele nanocompozite verzi sau ecologice

Aplicatiile biomedicale ale nanocompozitelor polimerice biodegradabile sunt ingineria tesuturilor [74], realizarea materialelor biocompatibile pentru implanturi [62], transportul medicamentelor in organism prin metoda incapsularii substantei farmacologic-actice in nanostructuri cu dimensiuni sub 100 nm [62

In ultimii ani au fost dezvoltate diferite tipuri de materiale intaritoare pentru nanocompozitele polimerice biodegradabile, cum ar fi

  • Nanoargile (silicati stratificati) [3,7]
  • Nanofibre celulozice [75]
  • Titanati ultra-fin stratificati [76]
  • Nanotuburi de carbon [77,78]

Un interes particular, deosebit, il reprezinta nanocompozitele polimerice bazate pe silicati stratificati modificati organic (OMLS) datorita proprietatilor superioare rasinilor polimerice nemodificate [61]. Aceste proprietati superioare sunt valabile in general pentru un continut redus de silicat (≤ 5 % procente masa). Pe baza tariei interactiilor dintre polimer si OMLS, sunt posibile doua structuri de nanocompozite, din punct de vedere termodinamic (fig. 7): nanocompozite intercalate si nanocompozite exfoliate. In cazul nanocompozitelor intercalate lanturile polimerice aflate in structura silicata sunt dispuse regulat din punct de vedere cristalografic indiferent de raportul dintre polimer si OMLS. In cazul nanocompozitelor exfoliate, straturile individuale de silicat sunt separate in matricea polimerica si se afla la anumite distante unele de altele in functie de continutul in OMLS al nanocompozitului.

INTERCALAT EXFOLIAT

 


Fig. 7 Doua tipuri de aranjamente posibile din punct de vedere termodinamic pentru un nanocompozit polimer/silicat

1.5. Metode de obtinere a nanocompozitelor polimerice pe baza de nanoargile.

O metoda foarte buna pentru obtinerea nanocompozitelor pe baza de nanoargile este intercalarea polimerilor in golurile din structura silicatului, folosind doua abordari

- inserarea unor monomeri potriviti in golurile dintre straturile de silicati urmata de polimerizare

- inserarea directa a lanturilor polimerilor in golurile silicatilor din solutie sau din topitura.

In ultimii ani intercalarea din topitura a devenit o metoda foarte atractiva pentru obtinerea nanocompozitelor polimer/OMLS deoarece aceasta metoda este compatibila cu tehnicile industriale moderne. Aceasta metoda presupune calirea materialului (incalzirea si racirea) static sau in cadrul operatiei de debitare (taiere, forfecare).

1.5.1. Metode Intercalative.

A-Inserarea lanturilor macromoleculare sau a unui prepolimer prin intercalarea din solutie in silicati stratificati.

B- Intercalarea din topitura.

1.5.2. Metode "In Situ" (cresterea lanturilor macromoleculare prin polimerizare in situ in silicati stratificati)

2. CORELAREA STRUCTURA-PROPRIETATI

Micromecanica teoretica actuala se bazeaza pe faptul ca proprietatile materialelor compozite, (ex. modulul de elasticitate Young) depind de proprietatile constituientilor, fractia volumica a componentelor, forma si aranjamentul incluziunilor, precum si de interfata matrice-incluziuni.

Aceasta teorie prevede faptul ca proprietatile materialelor compozite sunt independente de dimensiunile incluziunilor; acest lucru fiind in general valabil pentru sistemele in care au fost incluse particule/fibre cu dimensiuni micrometrice. Pentru sistemele nanocompozite nu este valabila aceasta teorie

Chan si colab. [54] au afirmat faptul ca anumite proprietati, cum ar fi modulul de elasticitate si fortele de tensiune scad in cazul nanocompozitelor cu matrice polipropilenica datorita modificarilor procesului de nucleatie, modificari determinate de nanoparticule (fig. 8).

Fig. 8 (a) PP pur; (b) PP continand 9,2 % filer (umplutura), procent volumic

Singh et al. [79] au studiat modificarile rezistentei la rupere a unei rasini poliesterice, modificari datorate adaugarii de particule de alumina cu dimensiunea de 20, 3,5 si 100 nm in diametru. Fig. 9 prezinta o crestere initiala a rezistentei la rupere, urmata de o scadere a acesteia la fractii volumice mari pentru particulele utilizate. Acest fenomen este atribuit aglomerarii nanoparticulelor in cazul maririi volumului de particule.

Fig. 9 Rezistenta la rupere (valorile normalizate) functie de fractia volumica procentuala a particulelor de aluminiu [79]

2.1. Caracterizarea fizico-chimica a materialelor

Metodele (fizice si sau chimice) de analiza

Morfologia la scala nanometrica deschide posibilitatea dezvoltarii unor modele de studiu interfaciale care permit evaluarea structurii si dinamicii catenelor inlantuite prin utilizarea tehnicilor conventionale macroscopice cum ar fi calorimetria cu scanare diferentiala, curentul stimulat termic, reometria, rezonanta magnetica nucleara sau diferitele metode spectroscopice [36,37,80,81].

Referinte

Suprakas SR, Masami O, "Polymer/layered silicate nanocomposites: a review from preparation to processing Prog. Polym. Sci. 28 (2003) 1539-1641.

Okada A, Kawasumi M, Usuki A, Kojima Y, Kurauchi T, Kamigaito O. "Synthesis and properties of nylon-6/clay hybrids" In: Schaefer DW, Mark JE, editors. "Polymer based molecular composites", MRS Symposium Proceedings, Pittsburgh, vol. 171; 1990. p. 45-50.

3. Giannelis EP, "Polymer layered silicate nanocomposites", Adv Mater 1996;8:29-35.

4. Giannelis EP, Krishnamoorti R, Manias E, "Polymer-silicate nanocomposites: model systems for confined polymers and polymer brushes", Adv Polym Sci 1999;138:107-47.

5. LeBaron PC, Wang Z, Pinnavaia TJ, "Polymer-layered silicate nanocomposites: an overview", Appl Clay Sci 1999;15:11-29.

6. Vaia RA, Price G, Ruth PN, Nguyen HT, Lichtenhan J, "Polymer/layered silicate nanocomposites as high performance ablative materials", Appl Clay Sci 1999;15:67-92.

7. Biswas M, Sinha RS, "Recent progress in synthesis and evaluation of polymer-montmorillonite nanocomposites", Adv Polym Sci 2001;155:167-221.

Giannelis EP, "Polymer-layered silicate nanocomposites:synthesis, properties and applications", Appl Organomet Chem 1998;12:675-80.

9. Xu R, Manias E, Snyder AJ, Runt J "New biomedical poly(urethane uera)-layered silicate nanocomposites", Macromolecules 2001;34:337-9.

10. Bharadwaj RK, "Modeling the barrier properties of polymerlayered silicate nanocomposites", Macromolecules 2001;34: 1989-92.

11. Messersmith PB, Giannelis EP, "Synthesis and barrier properties of poly(1 -caprolactone)-layered silicate nanocomposites", J Polym Sci, Part A: Polym Chem 1995;33:1047-57.

12 Yano K, Usuki A, Okada A, Kurauchi T, Kamigaito O, "Synthesis and properties of polyimide-clay hybrid", J Polym Sci, Part A: Polym Chem 1993;31:2493-8.

13. Kojima Y, Usuki A, Kawasumi M, Fukushima Y, Okada A, Kurauchi T, Kamigaito O, "Mechanical properties of nylon 6-clay hybrid", J Mater Res 1993;8:1179-84.

14. Gilman JW, Kashiwagi T, Lichtenhan JD, "Flammability studies of polymer-layered silicate nanocomposites", Sample J 1997;33:40-5.

15. Gilman JW, Flammability and thermal stability studies of polymer-layered silicate (clay) nanocomposites", Appl Clay Sci 1999;15:31-49.

16. Dabrowski F, Bras M Le, Bourbigot S, Gilman JW, Kashiwagi T, "PA-6 montmorillonite nanocomposite in intumescent fire retarded EVA", Proceedings of the Euro- fillers'99, Lyon-Villeurbanne, France; 6-9 September 1999.

17. Bourbigot S, LeBras M, Dabrowski F, Gilman JW, Kashiwagi T, "PA-6 clay nanocomposite hybrid as char forming agent in intumescent formulations", Fire Mater 2000;

18. Gilman JW, Jackson CL, Morgan AB, Harris Jr R, Manias E, Giannelis EP, Wuthenow M, Hilton D, Phillips SH, "Flammability properties of polymer-layered silicate nanocomposites, Propylene and polystyrene nanocomposites", Chem Mater 2000;12:1866-73.

19. Sinha RS, Yamada K, Okamoto M, Ueda K, "New polylactide/layered silicate nanocomposite: a novel biodegradable material", Nano Lett 2002;2:1093-6.

20. Dabrowski F, Bourbigot S, Delbel R, Bras ML, "Kinetic molding of the thermal degradation of polyamide-6 nanocomposite", Eur Polym J 2000;36:273-84.

21. Vaia RA, Jandt KD, Kramer EJ, Giannelis EP, "Kinetics of polymer melts intercalation. Macromolecules", 1995;28: 8080-5.

22. Vaia RA, Giannelis EP, "Lattice of polymer melt intercalation in organically-modified layered silicates", Macromolecules 1997;30:7990-9.

23. Vaia RA, Giannelis EP, "Polymer melts intercalation in organically-modified layered silicates: model predictions and experiment", Macromolecules 1997;30:8000-9.

24. Lee JY, Baljon ARC, Loring RF, Panagiopoulos AZ "Simulation of polymer melt intercalation in layered nanocomposites", J Chem Phys 1998;109:10321-30.

25. Balazs AC, Singh C, Zhulina E, "Modeling the interactions between the polymers and clay surfaces through self-consistent field theory", Macromolecules 1998;31: 8370-81.

26. Balazs AC, Singh C, Zhulina E, Lyatskaya Y, "Modeling the phase behavior of polymer-clay composites", Acc Chem Res 1999;32:651-7.

27. Fredrickson GH, Bicerano J, "Barrier properties of oriented disk composites", J Chem Phys 1999;110:2181-8.

28. Ginzburg VV, Balazs AC, "Calculating phase diagram of polymer-platelet mixtures using density functional theory: implication for polymer/clay composites", Macromolecules 1999;32:5681-8.

29. Baljon ARC, Lee JY, Loring RF, "Molecular view of polymer flows into a strongly attractive slit", J Chem Phys 2000;111: 9068-72.

30. Ginsburg VV, Singh C, Balazs AC, "Theoretical phase diagram of polymer/clay composites: the role of grafted organic modifier, Macromolecules 2000;33:1089-99.

31. Kuznetsov D, Balazs AC, "Scaling theory for end-functionalized polymers confined between two surfaces: predictions for fabricating polymer nanocomposites", J Chem Phys 2000; 112:4365-75.

32. Lee JY, Baljon ACR, Sogah DY, Loring RF, "Molecular dynamics study of intercalation of diblock copolymer into layered silicates", J Chem Phys 2000;112:9112-9.

33. Singh C, Balazs AC, "Effect of polymer architecture on the miscibility of polymer/clay mixtures", Polym Int 2000;49: 469-71.

34. Manias E, Chen E, Krishnamoorti R, Genzer J, Kramer EJ, Giannelis EP, "Intercalation kinetics of long polymers in 2 nm confinements", Macromolecules 2000;33:7955-66.

35. Ginsburg VV, Balazs AC, "Calculating phase diagrams for nanocomposites: the effect of adding end-functionalized chains to polymer/clay mixture", Adv Mater 2000;12: 1805-9.

36. Hackett E, Manias E, Giannelis EP, "Molecular dynamics simulations of organically modified layered silicates", J Chem Phys 1998;108:7410-5.

37. Hackett E, Manias E, Giannelis EP, "Computer simulation studies of PEO/layered silicate nanocomposites", Chem Mater 2000;12:2161-7.

38. Anastasiadis SH, Karatasos K, Vlachos G, Manias E, Giannelis EP, "Nanoscopic-confinement effects on local dynamics", Phys Rev Lett 2000;84:915-8. 1999;18:1761-3.

39. Zax DB, Yang DK, Santos RA, Hegmann H, Giannelis EP, Manias E, "Dynamical heterogeneity in nanoconfined poly(-styrene) chains", J Chem Phys 2000;112:2945-51.

40. Manias E, Kuppa V, "Molecular simulations of ultra-confined polymers. Polystyrene intercalated in layered-silicates. In: Vaia RA, Krishnamoorti R, editors. Polymer nanocomposites", ACS symposium series, vol. 804, Oxford: Oxford University Press; 2002. p. 193-207.

41. Kuppa V, Manias E, "Computer simulation of PEO/layeredsilicate nanocomposites: 2. Lithium dynamics in PEO/Li+ montmorillonite intercalates", Chem Mater 2002;14:2171-5.

42. Manias E, Kuppa V. Relaxation of polymers in 2-nm slitpores: confinement induced segmental dynamics and suppression of the glass transition. Colloids Surf., A 2001; 187-188:509-21.

Jordana J, Jacobb KI, Tannenbaumc R, Sharafb MA, Jasiukd I Experimental trends in polymer nanocomposites-a review Materials Science and Engineering A 393 (2005) 1-11.

44. Akita H, Hattori T, J. Polym. Sci. B: Polym. Phys. 37 (1999) 189-197.

45. Akita H, Kobayashi H, J. Polym. Sci. B: Polym. Phys. 37 (1999) 209-218.

46. Akita H, Kobayashi H, Hattori T, Kagawa K, J. Polym. Sci. B: Polym. Phys. 37 (1999) 199-207.

47. Chang J-H, An YU, J. Polym. Sci. B: Polym. Phys. 40 (2002) 670-677.

48. Zavyalov SA, Pivkina AN, Schoonman J, Solid State Ionics 147 (2002) 415-419.

B: Polym. Phys. 42 (2004) 1685-1693.

49. Siegel RW, Chang SK, Ash BJ, Stone J, Ajayan PM, Doremus RW, Schadler LS, Scr. Mater. 44 (2001) 2061-2064.

50. Wu CL, Zhang MQ, Rong MZ, Friedrich K, Compos. Sci. Technol. 62 (2002) 1327-1340.

51. Magaraphan R, Lilayuthalert W, Sirivat A, Schwank JW, Compos. Sci. Tech. 61 (2001) 1253-1264.

52. Vollenberg PHT, Heikens D, Polymer 30 (1989) 1656-1662.

53. Petrovic ZS, Javni I, Waddon A, Banhegyi G, J. Appl. Polym. Sci. 76 (2000) 133-151.

54. Chan C-M, Wu J, Li J-X, Cheung Y-K, Polymer 43 (2002) 2981-2992.

55. Rong MZ, Zhang MQ, Zheng YX, Zeng HM, Walter R, Friedrich K, Polymer 42 (2001) 167-183.

56. Yang F, Ou Y, Yu Z, J. Appl. Polym. Sci. 69 (1998) 355-361.

57. Ou Y, Yang F, Yu Z-Z, J. Polym. Sci. B: Polym. Phys. 36 (1998) 789-795.

58. Reynaud E, Jouen T, Gautheir C, Vigier G, Polymer 42 (2001) 8759-8768.

59. Li J-X, Wu J, Chan CM, Polymer 41 (2000) 6935-6937.

60. Park CI, Park OO, Lim JG, HJ Kim, Polymer 42 (2001) 7465-7475.

Suprakas SR, Mosto B, "Biodegradable polymers and their layered silicate nanocomposites: In greening the 21st century materials world Progress in Materials Science 50 (2005) 962-1079.

Orive G, Hernandez RM, Gascon AR, Dominguez-Gil A and Pedraz JL, "Drug delivery in biotechnology: present and future", Current Opinion in Biotechnology 2003, 14:659-664.

63. La Van D, Lynn DM, Langer R, "Moving smaller in drug discovery and delivery", Nat Rev Drug Discov 2002, 1:77-84.

64. La Van D, McGuire T, Langer R, "Small-scale systems for in vivo drug delivery", Nat Biotechnol 2003, 21:1184-1191.

American Ceramic Society (2004).

66. Thostenson ET, Li C, Chou T-W, Review, " Nanocomposites in context Composites Science and Technology 65 (2005) 491-516.

67. Merkulov VI, Lowndes DH, Wei YY, Eres G, Voelkl E, "Patterned growth of individual and multiple vertically aligned carbon nanofibers", Appl Phys Lett 2000;76(24):3555-7.

68. Endo M, Kim YA, Hayashi T, Fukai Y, Oshida K, Terrones M, et al., "Structural characterization of cup-stacked-type nanofibers with an entirely hollow core", Appl Phys Lett 2002;80(7):1267-9.

69. Endo M, Kim YA, Ezaka M, Osada K, Yanagisawa T, Hayashi T, et al., "Selective and eficient impregnation of metal nanoparticles on cup-stacked-type carbon nanofibers", Nano Letters 2003;3(6):723-6.

70. Ash BJ, Schadler LS, Siegel RW, Mater. Lett. 55 (2002) 83-87.

71. Shelley JS, Mather PT, DeVries KL, Polymer 42 (2001) 5849-5858.

72. Lindblad MS, Liu Y, Albertsson A-C, Ranucci E, Karlsson S, "Polymer from renewable resources", Adv Polym Sci 2002;157:139-61.

Sinha RS, Bousmina M, "Biodegradable polymers and their layered silicate nanocomposites: In greening the 21st century materials world", Progress in Materials Science 50 (2005) 962-1079.

74. Draye JP, Delaey B, Voorde AVD, Bulcke AVD, Rue BD, Schacht E, "In vitro and in vivo biocompatibility of dextran dialdehyde cross-linked gelatin hydrogel films", Biomaterials 1998;19:1677-87.

75. Mohanty AK, Drzal LT, Misra M, "Nano-reinforcement of bio-based polymers-the hope and reality", Polym Mater Sci Eng 2003;88:60-1.

76. Hiroi R, Sinha Ray S, Okamoto M, Shiroi T, "Organically modified layered titanate: a new nanofiller to improve the performance biodegradable polylactide", Macromol Rapid Commun 2004;25: 1359-63.

77. Mitchell CA, Bahr JL, Arepalli S, Tour JM, Krishnamoorti R, "Dispersion of functionalized carbon nanotubes in polystyrene", Macromolecules 2002;35:8825-30.

78. Andrews R, Wisenberger MC, "Carbon nanotube polymer composites", Curr Opin Solid State Mater Sci 2004;8:31-7.

Singh RP, Zhang M, Chan D, "Toughening of a brittle thermosetting polymer: e.ects of reinforcement particle size and volume fraction", J Mater Sci 2002;37(4):781-8.

80. VanderHart DL, Asano A, Gilman JW. NMR measurements related to clay dispersion quality and organic-modifier stability in nylon 6/clay nanocomposites. Macromolecule 2001;38:3819-22.

81. Loo LS, Gleason KK. Fourier transforms infrared investigation of the deformation behavior of montmorillonite in nylon 6/nanoclay nanocomposites. Macromolecules 2003;36:2587-90.

2. CORELAREA STRUCTURA-PROPRIETATI

2.1. CARACTERIZAREA FIZICO-CHIMICA A MATERIALELOR

Metodele (fizice si sau chimice) de analiza

Nanocompozitele se compun dintr-un polimer si un silicat stratificat (modificat sau nu) frecvent prezentand proprietati mecanice si de material imbunatatite remarcabil cand le comparam cu cele ale polimerilor clasici continand o mica cantitate (£ 5%) de silicat stratificat. Imbunatatirile includ un coeficient mai mare, crescand durabilitatea si rezistenta termica, scazand permeabilitatea gazului si inflamabilitatea , si crescand biodegradabilitatea a polimerilor biodegradabili.

Fig. 66. Imagini in camp stralucitor TEM ale (a) PLACN12 (C8, 1,7%) si (b) PLACN2 (C16, 3%)

Motivul principal al acestor proprietati imbunatatite din nanocompozite este interactiunea interfaciala mai puternica dintre matrice si silicatul stratificat, comparat cu sistemele conventionale ranforsate cu umplutura.

2.1.1.Proprietati mecanice.

2.1.1.1.Analize mecanice dinamice.

Analizele mecanice dinamice (DMA) masoara raspunsul dat de un material la o deformatie oscilatorie (aici in modul tensiune-torsiune) ca o functie de temperatura. Rezultatele DMA sunt compuse din trei parametrii: (a) coeficientul de stocare (GI); (b) coeficientul de pierdere (GII), si (c) tang d, raportul (GII/ GI), folositor pentru determinarea pozitiei tranzitiilor mobilitatii moleculare, cum ar fi temperatura tranzitiei sticloase (Tg).

DMA au fost folosite pentru a studia dependenta temperaturii de coeficientul de stocare al PMMA pe pentru matia de nanocompozit sub conditii experimentale diferite Fig. 71 aratand dependenta temperaturii de GI si tang d, pentru variate nanocompozite bazate pe copolimeri, si corespunzator amestecurilor fara argila de copolimer/QA+. Masuratorile au fost realizate folosind un instrument RDAII cu frecventa oscilatorie de 6,28 rad/s, cu o amplitudine de 0,05%, si o rata de incalzire de 20C/min.

Fig. 67. Exemple WAXD de smectite, MMT, si mica nanocompozite cu C16 organo-argila si acelasi continut de argila

Fig. 68. Prezentarea schematica a straturilor de silicat in OMLS si in diferite nanocompozite

Fig. 69. C18-MMT, diferiti PBSCNs Fig. 70. Imagini TEM de PBSCN cu 2,8 % MMT

si PBS pur. Linia punctata indica locul

silicatului (001) reflectat de C18-MMT

dispersat in matricea PBS

Pentru amestecurile PMMA/SPN10 si PMMA/QA+, nu este nici o diferenta intre dependenta de temperatura de GI sau tang d (vezi Fig. 71a). Pana la incorporarea de AEA, PAA, si AA pana la PMMA, o crestere puternica in GI se dezvolta la toate temperaturile in timp ce matricea copolimera apare. Acest comportament indica ca adaugarea de copolimer are un efect puternic asupra proprietatilor elastice ale matricelor corespunzatoare. Pe de alta parte, peak-urile tang d ale nanocompozitelor sunt indreptate catre temperaturi mai mici decat pentru amestecurile fara argila corespunzatoare. Comparate cu alte sisteme, comportamentul ui PMMA-AA(1)/SPN10 este putin diferit; peste Tg, GI este aproape acelasi pentru nanocompozite si matricele fara argila corespunzatoare, desi o extindere si o mica schimbare de temperatura la tang d este observata. Comportamentul poate fi facut la intercalarea lanturilor copolimerice in galeriile straturilor de argila, care conduc la impiedicarea mobilitatii ale segmentelor copolimrului aproape de interfata. Aceasta presupunere a fost sprijinita de modelele WAXD ale acestor nanocompozite (vezi Fig. 21d), care prezinta slab, dar peak-uri semnificative din planurile expandate (002) si (003).

Fig. 71. Dependenta temperaturii de GI si tang d pentru nanocompozite si amestecurilor copolymer/QA+ corespunzatoare fara argila

Fig. 72. Dependenta temperaturii de GI; GII si tang d pentru matricea PP-MA

si diferite PPCNs

Fig. 72 arata dependenta temperaturii de GI, GII si tang d de diferite PPCNs si matricea corespunzatoare PP-MA. Pentru toate PPCNs, este o puternica intensificare a modulilor peste nivelul temperaturii investigate, care indica raspunsurile plastice si elastice ale PP fata de deformatie sunt puternic influentate in prezenta de OMLS. Sub Tg, intensificarea lui GI este clar in PPCNs intercalat.

In contrast, curbele tang d pentru PPCNs arata doua peak-uri: unul la -50C si altul peak clar intre 50 si 900C. McCrum si colaboratorii au demonstrat ca curba tang d a PP trei exponati relaxati localizati in vecinatatea - 800C (g), 100C (Tg) si 1000C (a). Relaxarea dominanta la circa 100C este relaxarea sticla-cauciuc a portiunii amorfe de PP.

Fig. 73. Spectre ale dinamicii mecanice: (a) modulul de stocare, EI; (b) modulul de pierdere, EII; si (c) factorul de pierdere tang d ca o functie de temperature pentru PP pur si PPCN

Proprietatile mecanice dinamice ale PP si PPCNs pure preparate cu EM-MMT sunt aratate in Fig. 73 [304]. Aceste rezultate arata clar ca incorporarea de EM-MMT in matricea PP reprezinta o remarcabila crestere in rigiditate si o descrestere in tang d. Curbele GI arata un palier elastic largit, aratand ca adaugarea de EMMMT induce un efect de intarire; la temperaturi foarte mari acest efect de intarire a rigiditatii. Acest comportament in plus indica sporirea stabilitatii termo-mecanice a acestor materiale la temperaturi inalte.

Similar cu sistemele anterioare, acesti PPCNs arata doua peak-uri in tang d (vezi Fig. 73c). Procesul dominant de relaxare la circa 100C este relaxarea sticlos-elastica a portiunii amorfe de PP. Aparitia peak-ului slab sub pentru ma de umar la circa 1000C este asociat cu regiuni (a-faze) cristaline ale PP. Un alt fenomen interesant este acela ca valorile Tg ale PPCNs nu descresc mai mult peste un continut de EM-MMT de 3%. Oricum, mecanismul pentru acest comportament ne ste inca inteles.

Imbunatatirea remarcabila in GI; in plus puternica interactie dintre matricea de OMLS, este observata clar in nanocompozitele N6/OMLS

Fig. 74 reprezinta dependenta temperaturii de GI, GII si tang d a matricei N6 si diferite nanocompozite (N6CNs). Detaliile ale dinamicii temperaturii din testul de rampa pentru N6 pur si diferite N6CNs sunt sumarizate in Tabelul 12. Toate N6CNs arata o mare crestere in modul la toate temperaturile.

Cresterile in GI observate la dimensiunea particulelor de argila dispersate este in plus demonstrata in PLACNs . In ordine pentru a intelege efectul compatibilizarilor peste proprietatile morfologice si mecanice, autorii deasemenea au preparat PLACNs cu foarte mici cantitati de oligo(ε -caprolactona) (a-PCL).

Detaliile si denumirile compozitiei pentru diferite tipuri de nanocompozite sunt prezentate in Tabelul 8

Fig. 74. Dependenta temperaturii de GI; GII si tang d pentru matricea N6 si diferiti N6CNs.

Tabelul 12

Testele sumare de DMA pentru N6 si diferiti N6CNs sub diferite niveluri de temperatura

Partile a si b ale Fig. 75 arata dependenta temperaturii de GI; GII si tang d a matricii de PLA si respectiv a diferitelor PLACNs. Pentru toate PLACNs, sporirea de GI poate fi vazuta peste nivelul de temperatura investigat cand comparat cu matricea, indica ca C18-MMT are un efect puternic asupra proprietatilor elastice ale PLA pur. Sub Tg, sporirea de GI este clara pentru diferite intercalate PLACNs. Pe de alta parte, toate PLACNs arata o mai mare crestere in la temperaturi mari comparat fata de cele ale matricelor de PLA.

Fig. 75. Dependenta de temperatura a GI; GII si nivelul lor tang d pentru PLACNs si matricele corespunzatoare: (a) fara o-PCL si (b) cu o-PCL

Acest lucru este facut celor doua intariri mecanice de catre particulele de argila si intercalare extinsa la temperaturi mari . Peste Tg, cand materialele devin moi, efectul de intarire ale particulelor de argila devin proeminente fata de miscarea restrictiva a lanturilor de polimer. Acesta este acompaniat de observarea cresterii de GI.

La cealalta extrema, probele PLACN4 si PLACN5 arata cresteri mari ale GI comparate cu cele ale probei de PLACN2 cu comparabila argila adaugata, si fata de matricele PLA/o-PCL continand pana la 0,5% o-PCL (vezi Tabelul 9 Fig. 75b). prezenta unor mici cantitati de o-PCL nu conduc la o mare schimbare sau largire a curbelor tang d Oricum, o mare crestere in GI peste Tg devine evidenta, indicand ca marea anisotropie ale particulelor floculate dispersate sporeste componentul tinta. Valorile lui GI la diferite niveluri de temperatura ale diferitelor PLACNs si matricele corespunzatoare fara argila sunt reprezentate in Tabelul 13. PLACNs cu o foarte mica cantitate de o-PCL (PLACN4 si PLACN5) expun o foarte mare sporire a proprietatilor mecanica comparate cu cele ale PLACN cu comparabila argila adaugata (PLACN2). Factorul esential care guverneaza sporirea proprietatilor mecanice in nanocompozite este nivelul aspectului al particulelor de argila disperate

Tabelul 13

Valoarea GI a diferite PLACNs si matricele corespunzatoare fara argila la diferite niveluri de temperatura

Din figurile TEM (vezi Fig. 62), este clar vazut ca, in prezenta unei foarte mici cantitati de o-PCL, floculatia particulelor de argila dispersate au loc, iarasi datorita puternicelor interactiuni margine-margine ale particulelor de argila. Nivelele aspectului 2D ale particulelor de argila dispersate Lclay/dclay estimate din observarea TEM sunt 22 pentru PLACN4 si 12 pentru PLACN2 (vezi Tabelul 9). Acest mare nivel de aspect conduce la sporirea observata a proprietatilor mecanice.

Tabelul 8

Compozitia si parametrii caracteristici ai diferitelor PLACNs bazate pe PLA, o-PCL si C18-MMT

Tabelul 9

Comparatie intre factorii de pentru mare a PLACN2 si PLACN4 obtinuta din structurileWAXD si obervatiile TEM

Fig. 62. Imagini campuri stralucitoare TEM: (a) PLACN2 ( 10,000), (b) PLACN4 ( 10,000), (c) PLACN2 ( 400,000), si (d) PLACN4 ( 40,000). Entitatile intunecate sunt sectiunile de trecere a OMLS intercalat, si zonele luminoase sunt matricele

Ipoteza ca o crestere in GI depinde direct de nivelul aspectului al particulelor de argila dispersate este de asemenea clar observate in PBSCNs. Dependenta temperaturii de GI pentru PBS si diferiti PBSCNs sunt prezentate in Fig. 76a. Natura cresterii lui GI in PBSCNs cu temperatura este cumva diferita fata de teoriile bina stabilite, care explica comportamentul similar observat in sistemele care sunt fie intercalate (PP-MA/MMT) sau exfoliate (N6/MMT) . In sistemul precedent, GI creste tipic cu circa 40-50°%, cand comporat cu cele ale matricii mult sub Tg, in timp ce peste Tg, este o mare crestere (>200%) in GI. Acest comportament este comun pentru nanocompozitele prezntate mai sus, si motivul a fost aratat a fi efectul puternic de intarire al particulelor de argila peste Tg (cand materialele devin moi). Oricum, in cazul PBSCNs, ordinea cresterii in GI este aproape aceiasi sub si peste Tg, si acest comportament poate avea loc la un extreme de mic Tg (-290C) al matricii de PBS.

Peste nivelul de temperatura de - 50 pana la -100C, cresterea in GI este de 18% pentru PBSCN1, 31% pentru PBSCN2, 67% pentru PBSCN3 si 167% pentru PBSCN4 comparate cu cele ale PBS pur. In plus, la temperatura camerei PBSCN3 si PBSCN4 arata o mare crestere in GI, 82 si 248%, respectiv, comparate cu cele de PBS pur, in timp ce cele de PBSCN1 si PBSCN2 sunt mai mari cu 18,5 si 44% La 900C. Doar PBSCN4 prezinta o foarte mare crestere a GI comparat cu celelalte trei PBSCNs.

In contrast, PBSCNs preparat cu qC16-sap prezinta o relativa mica crestere a GI comparativ cu cele de PBSCNs preparate cu C18-MMT (vezi Fig. 76b). Pentru PBSCN6, crestereea in GI este de 102,5% la -50 0C, 128,6% la 25 0C si 100% la 90 0C, comparat cu PBS. Aceste valori sunt mult mai mici comparate cu cele de PBSCN4, desi ambele specii contin un comparabil continut in argila (parte anorganica). Doi factori pot fi sugerati pentru o foarte mare crestere a modulului in cazul PBSCN4, comparat cu cel de PBSCN6, unul din factori este foarte marele format al aspectului al particulelor de argila dispersate si alt factor este bine ordonata structura intercalata in PBSCN4.

Fig. 76. Dependenta temperaturii GI; GII si formatul lor tang d pentru (a) PBSCNs (preparat cu C18-MMT) si PBS pur, (b) PBSCNs

Fig. 77. Grafice ale GInanocomposite/GImatrix vs. vol% de argila pentru nanocompozite diferite. Coeficientul Einstein kE este aratat cu numar in chenar, liniile arata rezultatele calculate de teoria lui Halpin si Tai's cu variate kE.

Dependenta continutului de argila de GI pentru diferite tipuri de nanocompozite obtinute mult sub Tg sunt prezentate in Fig. 77, aratand coeficientul Einstein, kE; care este derivat folosind expresia teoretica a lui Halpin si Tai's modificata de Nielsen si repezinta formatul aspectului (L/D) al particulelor dispersate de argila fara intercalare. expresia Haplin-Tai's-Nielsen al modulului nanocompozitelor, GInanocomposite este dat de

Aici, GImatrix si GIclay sunt module de stocare ale matricei (aici PLA, PBS, PP-MA si N6) si respectiv argila. X este o constanta care depinde de tipul structurii nanocompozitului, este legata de formatul aspectului, si ψclay si ψm sunt fractiuni de volum al argilei intarite si respectiv maximul fractiuni de volum de umplere (in general egal cu 0,63). Presupunerea ca valoarea lui GIclay este egala cu 170 GPa , permitand estimarea dependentei compozitiei de GInanocomposite/GImatrix folosind ecuatia de mai sus. In plus, valorile lui kE au fost estimate prin selectarea unei valori apropiate din cele mai bune potrivite pentru experiment obtinand graficele GInanocomposite/GImatrix vs. ψclay (vezi Fig. 77

Din Fig. 77, este clar observat ca PBSCNs prezinta o mare crestere in GI comparat cu al altor nanocompozite avand acelasi continut de argila in matrice. PPCNs sunt bine cunoscute ca sisteme intercalate, N6CNs sunt bine stabilite ca nanocompozite exfoliate, PLACNs sunt considerate nanocompozite intercalate-si-floculate, in timp ce PBSCNs sunt nanocompozite intercalate-si-extins floculate . Datorita interactiunii puternice dintre gruparilor hidroxilate capat-capat, particulele de argila sunt uneori floculate in matricea de polimer. Ca rezultat al acestei floculatii, lungimea particulelor de argila creste enorm, rezultand intr-o crestere corespunzatoare peste toate formatului aspectului. Pentru prepararea unei mase moleculare mari PBS, grupari finale diisocianat sunt in general folosite ca o prelungire de lant . Aceste grupari finale de isocianat fac legaturi uretanice cu cu cele de hidroxi terminale LMW PBS. Fiecare masa moleculara mare a lantului PBS doua din acest tip de legaturi, si de aici (vezi ilustrarea schematica in Fig. 78

Fig. 78. Formarea legaturilor uretanice in mase moleculare mari PBS

Fig. 79. Formarea legaturilor de hidrogen intre PBS si argila, care conduce la floculatia silicatilor stratificati dispersati

Aceste legaturi tip uretanive conduc la o puternica interactie cu suprafata silicata prin formarea legaturilor de hidrogen, si de aici floculatie puternica (vezi Fig. 79). Pentru acest motiv, formatul aspectului al particulelor de argila dispersate este mult mai mare in cazul lui PBSCNs comparat cu alte nanocompozite, si de aici marea intensificare a modulului.

Fig. 80. Ilustrarea schematica pentru formarea legaturilor de hidrogen in nanocompozite

N6/MMT.

2.1.2. Proprietatile de intindere

Modulul de intindere al unui material polimeric a fost prezentat a fi remarcabil imbunatatit cand nanocompozitele sunt formate cu silicati stratificati. Nanocompozitele N6 preparate prin inele intercalate in situ deschizand polimerizarea ε - caprolactama, conducand la formarea unor nanocompozite exfoliate, prezinta o crestere drastica in prorietatile de intindere la mai degraba mici continuturi de umplutura. Motivul principal pentru drastica imbunatatire in modulul de intindere in nanocompozitele N6 este puternica interactiune dintre matrice si silicati stratificati prin formarea legaturilor de hidrogen, cum este aratat in Fig. 80

In cazul nanocompozitelor, intinderea imbunatatita a modulului depinde direct pe media lungimii particulelor de argila dispersate, si de aici formatul aspectului. . Fig. 81 reprezinta dependenta modulului de intindere E masurat la 1200C pentru nanocompozite exfoliate N6 cu diferite continuturi de argila, obtinute prin polimerizarea intercalata in situ a ε-caprolactama in prezenta de acid aminododecanoic protonat -MMT modificat si saponit. Mai mult, diferenta in exfolierea intinsa, cum se observa pentru nanocompozite bazate pe N6 sintetizate prin polimerizarea intercalata in situ a ε-caprolactama folosind Na+ - MMT si diferiti acizi, influentat puternic modulul nanocompozitelor.

Tabelul 14

Proprietatile mecanice ale IpotNCH sintetizat in prezenta acidului fosoric si prin metoda intercaltiva in situ.

Tabelul 14 prezinta modulul de intindere al 1potNCH impreuna cu N6 pur si NCH preparate prin polimerizare in situ intercalata cu deschidere de inel a ε-caprolactama . Modulul excelent in cazul 1potNCH este atribuita dispersarii uniforme a silicatilor stratificat. In plus, 1potNCH si-a imbunatatit proprietatile mecanice cand comparam cu NCH.

Tabelul 14 summarizes the tensile modulus of

1potNCH together with neat N6 si NCH prepared

via in situ intercalative ring-opening polymerization

of1 -caprolactam . The excellent modulus in the

case of 1potNCH is attributed to the uniformly

dispersed silicate layers. Furthermore, 1potNCH has

improved mechanical properties when compared with

NCH. The polymer matrix in the nanocomposites

prepared by a one pot synthesis is the homopolymer of

N6, whereas in the case of NCH prepared via

intercalative ring-opening polymerization, the matrix

is a copolymer of N6 si a small amount of N12. The

presence of N12 may give rise to the lower modulus.

One can observe variations of the modulus of the

nanocomposites based on the various kinds of

acids used to catalyze the polymerization

The WAXD peak intensity (Im; inversely related to

the exfoliation of clay particles) also depends on the

nature of the acid used to catalyze the polymerization

process. pentru   an increase in the Im values, a parallel

decrease in the modulus is observed, indicating that

exfoliated layers are the main factor responsible pentru

the stiffness improvement. Intercalated particles,

having a less important aspect ratio, play a minor

role. These observations are further confirmed in

Fig. 82, which presents the evolution of the tensile

modulus at room temperature of N6 nanocomposites

obtained by melt extrusion as a function of the filler

content

The effect of MMT content si N6 molecular

weight on the tensile modulus of nanocomposites

prepared using MMT modified with (HE)2M1R1 is

shown in Fig. 83. The addition of organoclay leads to

a substantial improvement in stiffness pentru the

composites based on each of the three N6, i.e.

LMW, MMW si HMW. Interestingly, the stiffness

increases with increasing matrix molecular weight at

any given concentration even though the moduli of the

neat N6 are all quite similar. Tabelul 15 summarizes the

moduli si other mechanical properties of the virgin

Fig. 81. Effect of clay content on tensile modulus in case of

N6/OMLS nanocomposites prepared via melt extrusion

Reproduced from Kojima, Usuki, Kawasumi, Okada, Kurauchi and

Kamigaito by permission of Materials Research Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1598

materials si selected (HE)2M1R1 organoclay/N6

nanocomposites. The slightly larger modulus of

2.82 GPa pentru   LMW may be the result of a higher

degree of crystallinity, creating faster crystallization

kinetics during the cooling of the specimen during

injection molding.

Similar trends with respect to the level of

organoclay content si molecular weight are evident

in the yield strength results. The dependence of yield

strength on MMT content si molecular weight is

shown in Fig. 84. Yield strength increases with the

content of MMT, however, there are notabelul differences

in the level of strength improvement pentru pure

polyamides. The HMW- si MMW-based nanocomposites

show a steady increase in strength with content

of clay, while the LMW-based nanocomposites show

a less pronounced effect. The differences in strength

improvement with respect to molecular weight are

very prominent at the highest clay content. The

increase in strength relative to the virgin matrix pentru the

HMW composite is nearly double to that of the LMW

composite.

The relationship between MMT content and

elongation at break pentru   the different matrices is

shown in Fig. 85 pentru two different rates of extension.

Fig. 85a shows that the virgin polyamides are very

ductile at a test rate of 0.51 cm/min. With increasing

clay content the ductility gradually decreases, however,

the HMW si MMW based composites attain

reasonable levels of ductility at MMT concentrations

as high as 3.5 wt%. The elongation at break pentru the

LMW-based nanocomposites decreases rapidly at low

MMT content (around 1 wt%). The larger reduction in

the LMW-based systems may be due to the presence

of stacked silicate layers, as seen in TEM photographs

(see Fig. 46). In contrast, the higher testing rate of

5.1 cm/min yields similar trends, as shown in Fig. 85b

but the absolute level of the elongation at break is

significantly lower. Interestingly, the strain at break

for LMW composites is relatively independent of the

rate of extension, similar to what has been observed in

glass fiber reinforced composites. Even at the highest

clay content, the HMW composite exhibits ductile

fracture, whereas the LMW- si MMW-based

nanocomposites fracture in a brittle manner at the

highest clay content.

In the case of PPCNs, most studies report the

tensile properties as a function of clay content. The

results of an Instron study of a neat-PP/f-MMT

composite compared to a PP/2C18-MMT 'conventional'

composite are shown in Fig. 86. In PP/layered

silicate nanocomposites, there is a sharp increase in

tensile modulus pentru   very small clay loading

(#3 wt%), followed by a much slower increase

beyond a clay loading of 4 wt%. This is behavior

characteristic of PLS nanocomposites. With an

increase in clay content, strength does not change

markedly compared to the neat-PP value, si there is

only a small decrease in the maximum strain at break.

Conventional composites of PP with the same fillers

do not exhibit as much of an improvement in their

tensile modulus. On the other hand, as the PP/layered

silicate interaction is improved, pentru example when

Fig. 82. Dependence of tensile modulus ðEÞ on clay content

measured at 120 8C . Reproduced from Alexander si Doubis

by permission of Elsevier Science Ltd, UK.

Fig. 83. Effect of MMT content on tensile modulus pentru LMW,

MMW, si HMW based nanocomposites . Reproduced from

Fornes, Yoon, Keskkula si Paul by permission of Elsevier Science

Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1599

MA functional groups are incorporated into the

polymer, the stress is much more efficiently transferred

from the polymer matrix to the inorganic filler,

resulting in a higher increase in tensile properties.

Parts a si b of Fig. 87 represent the dependence of

the tensile modulus si strength on MMT content of

various PPCNs prepared by melt extrusion of PP-MA

and C18-MMT, respectively. The modulus of the

PPCNs systematically increases with increasing clay

content, as does the tensile strength up to 4 wt%,

where it levels off.

If the interaction between nanocomposite components

is not thermodynamically favorable, these

properties will change during processing because the

nanocomposite structure will change. Recent work by

Reichert et al. systematically found the

dependencies on compatibilizer functionality and

organic modification, si revealed that considerable

tensile property enhancement could be achieved only

when appropriate PP-MA compatibilizers were used

to pretreat the OMLS in conjugation with specific

organic modifications of the MMT. Similar materials

under different processing conditions showed much

smaller improvements in the practical material

properties

The tensile properties of various PPCNs prepared

with EM-MMT, a new type of co-intercalated MMT,

are summarized in Fig. 88 [304]. The PPCN containing

1 wt% EM-MMT is abbreviated as PPCN1, while

the PPCNs with 3, 5 si 7 wt% of EM-MMT are

abbreviated as PPCN3, PPCN5, si PPCN7, respectively.

The tensile strength of the PPCNs increase

rapidly with increasing EM-MMT content from 0 to

5 wt%, but the trend is less pronounced when the clay

content increases beyond 5 wt%. A similar trend is

observed pentru   the tensile modulus. In contrast, the

notched Izod impact strength of the PPCNs is

Tabelul 15

Mechanical properties of some N6/(HE)2M1R1 nanocomposites

N6/(HE)2M1R1 nanocomposites Modulus

(GPa)

Yield strength

(MPa)

Straina

Elongation at break

(%) crosshead speed

Izod impact strength

(J/m)

0.51 cm/min 5.1 cm/min

LMW

0.0 wt% MMT 2.82 69.2 4.0 232 28 36.0

3.2 wt% MMT 3.65 78.9 3.5 12 11 32.3

6.4 wt% MMT 4.92 83.6 2.2 2.4 4.8 32.0

MMW

0.0 wt% MMT 2.71 70.2 4.0 269 101 39.3

3.1 wt% MMT 3.66 86.6 3.5 81 18 38.3

7.1 wt% MMT 5.61 95.2 2.4 2.5 5 39.3

HMW

0.0 wt% MMT 2.75 69.7 4.0 3.4 129 43.9

3.2 wt% MMT 3.92 84.9 3.3 119 27 44.7

7.2 wt% MMT 5.70 97.6 2.6 4.1 6.1 46.2

Reproduced from pentru nes, Yoon, Keskkula si Paul by permission of Elsevier Science Ltd, UK.

a Strain at yield point measured during modulus si yield strength testing using a crosshead speed of 0.51 cm/min.

Fig. 84. Effect of MMT content on yield strength pentru LMW,

MMW, si HMW based nanocomposites . Reproduced

from pentru nes, Yoon, Keskkula si Paul by permission of Elsevier

Science Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1600

constant, within experimental error, in the EM-MMT

content range between 0 si 7 wt%.

The evolution of the tensile modulus pentru the epoxy

matrix with three different types of layered silicates is

presented in Fig. 89 [228]. A C18-MMT, a C18Amagadiite,

and a magadiite modified with methyloctadecylammonium

cation (C18A1M-magadiite)

were used pentru   nanocomposite preparation. This figure

shows a significant increase in the modulus pentru the

MMT-based nanocomposites with filler content of

4 wt%. According to the present authors, this behavior

is due to the difference in layer charge between

magadiite si MMT. Organomagadiites have a higher

layer charge density, si subsequently higher alkylammonium

content than organo-MMT. As the

alkylammonium ions interact with the epoxy resin

while polymerizing, dangling chains are pentru med, and

more of these chains are pentru med in the presence of

organomagadiites. These dangling chains are known

to weaken the polymer matrix by reducing the degree

of network cross-linking, then compromising the

reinforcement effect of the silicate layer exfoliation.

For thermoset matrices, a significant enhancement

in the tensile modulus is observed pentru an exfoliated

structure when alkylammonium cations with different

chain length modified MMTs were used pentru nanocomposite

preparations, with the exception of the

MMT modified with butylammonium, which only

gives an intercalated structure with a low tensile

modulus.

In a recent work, Mulhaupt et al. reported the

correlations between polymer morphology, silicate

structure, stiffness, si toughness of thermoset

Fig. 85. Effect of MMT content on elongation at break pentru LMW, MMW, si HMW based nanocomposites at a crosshead speed of

(a) 0.51 cm/min si (b) 5.1 cm/min Reproduced from pentru nes, Yoon, Keskkula si Paul by permission of Elsevier Science Ltd, UK.

Fig. 86. Tensile characterization of the PP/f-MMT nanocomposites

(B) by Instron. pentru   comparison, conventionally filled PP/2C18-

MMT 'macro' composites are also shown (W) . Reproduced

from Manias, Touny, Strawhecker, Lu si Chang by permission of

American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1601

nanocomposites as a function of layered silicate type

and content. They suggest that the main factor pentru the

matrix stiffness improvement resides in the pentru mation

of supramolecular assemblies obtained by the presence

of dispersed anisotropic laminated nanoparticles.

They also describe a stiffening effect when the

MMT is modified by a functionalized organic cation

(carboxylic acid or hydroxyl groups) that can strongly

interact with the matrix during curing.

The tensile properties of APES/Cloisite 30B and

APES/Cloisite 10A nanocomposites at various clay

contents are presented in Tabelul 16 [398]. In

comparison to the APES, the tensile strength and

modulus have been improved with a slight decrease in

elongation at break. APES/Cloisite 30B nanocomposites

exhibit a much higher tensile strength and

modulus compared to the APES/Cloisite 10A nanocomposites.

This is also attributed to the strong

interaction between the APES matrix si Cloisite

30B. These results further confirm the importance of

strong interaction between matrix si clay, which

ultimately leads to better overall dispersion, as

already observed by TEM analysis.

4.1.3. Flexural properties

Nanocomposite researchers are generally interested

in the tensile properties of final materials, but

there are very few reports concerning the flexural

properties of neat polymer si its nanocomposites

with OMLS. Very recently, Sinha Ray et al.

reported the detailed measurement of flexural properties

of neat PLA si various PLACNs. They

conducted flexural property measurements with

injection-molded samples according to the ASTM

D-790 method. Tabelul 17 summarizes the flexural

modulus, flexural strength, si distortion at break of

neat PLA si various PLACNs measured at 25 8C.

There was a significant increase in flexural modulus

for PLACN4 when compared to that of neat PLA,

followed by a much slower increase with increasing

OMLS content, si a maximum at 21% pentru PLACN7.

Fig. 88. Effect of clay loading on: (a) tensile modulus, si (b) tensile strength of PPCNs . Reproduced from Liu si Wu by permission of

Elsevier Science Ltd, UK.

Fig. 87. Relative moduli of various PP-based nanocomposites, each

normalized by modulus of the respective neat PP. (a) PP-based

nanocomposites with: f-MMT (B), C18-MMT (K), si 2C18-MMT

(W). (b) PP-g-MA/2C18-MMT nanocomposite (B) si PP hybrids

with various PP-g-MA pretreated organically modified MMT: C18-

MMT right triangle open), C18-MMT (W, K), si C8-MMT (K, K)

. Reproduced from Manias, Touny, Strawhecker, Lu and

Chang by permission of American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1602

On the other hand, the flexural strength si distortion

at break shows a remarkable increase with PLACN4,

then gradually decreases with OMLS loading.

According to the author, this behavior may be due

to the high OMLS content, which leads to brittleness

in the material.

4.2. Heat distortion temperature

Heat distortion temperature (HDT) of a polymeric

material is an index of heat resistance towards applied

load. Most of the nanocomposite studies report HDT

as a function of clay content, characterized by the

procedure given in ASTM D-648. Kojima et al.

first showed that the HDT of pure N6 increases up to

90 8C after nanocomposite preparation with OMLS.

In their further work they reported the clay

content dependence of HDT of N6/MMT nanocomposites.

In N6/MMT nanocomposites, there is a

marked increase in HDT from 65 8C pentru the neat N6

to 152 8C pentru   4.7 wt% nanocomposite. Beyond that

wt% of MMT, the HDT of the nanocomposites level

off. They also conducted HDT tests on various N6

nanocomposites prepared with different lengths of

clay si found that the HDT also depends on the

aspect ratio of dispersed clay particles . Like all

other mechanical properties, the HDT of 1potNCH is

higher than that of NCH prepared by in situ

intercalative polymerization

Since the degree of crystallinity of N6 nanocomposites

is independent of the amount si nature of

clay, the HDT of N6 nanocomposites is due to the

presence of strong hydrogen bonds between the

matrix si silicate surface (see Fig. 80). Although

N6 in nanocomposites results in a different crystal

phase (g-phase) than that found in pure N6, this

different crystal phase is not responsible pentru the higher

mechanical properties of N6 nanocomposites because

the g-phase is a very soft crystal phase. The increased

mechanical properties of N6 nanocomposites with

increasing clay content is due to the mechanical

reinforcement effect.

The nano-dispersion of MMT in the PP matrix

also promotes a higher HDT . The HDT of PP

and its nanocomposites based on f-MMT and

Tabelul 16

Tensile properties of APES/Closite 30B nanocomposites

Closite 30B content

(wt%)

Modulus

(kgf/cm2)

Strength

(kgf/cm2)

Elongation at break

Reproduced from Lee, Park, Lim, Kang, Li, Cho si Ha by

permission of Elsevier Science Ltd, UK.

Fig. 89. A comparison of (A) the tensile strengths si (B) tensile moduli pentru   epoxy nanocomposites prepared from C18A-MMT, C18Amagnitide,

and C18A1M-magnitide. The silicate loading was determined by calcining the composites in air at 650 8C pentru   4 h at a heating rate of

2 8C/min . Reproduced from Wang, Lan si Pinnavaia by permission of American Chemical Society, USA.

Tabelul 17

Comparison of materials properties between neat PLA si various

PLACNs prepared with octadecyltrimethylammonium modified

MMT

Materials properties PLA PLACN4 PLACN5 PLACN7

Bending modulus (GPa) 4.8 5.5 5.6 5.8

Bending strength (MPa) 86 134 122 105

Distortion at break (%) 1.9 3.1 2.6 2

Reproduced from Sinha Ray, Yamada, Okamoto si Ueda by

permission of Elsevier Science Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1603

alkylammonium modified MMT are summarized in

Tabelul 18. Like previous systems, there is a

significant increase in HDT, from 109 8C pentru the

neat PP to 152 8C pentru   a 6 wt% of clay, after which

the HDT of the nanocomposites levels off. This

improvement in HDT pentru   neat PP after nanocomposite

preparation originates from the greater mechanical

stability of the nanocomposite as compared to

neat PP, since there is no increase in melting point

of neat PP after nanocomposite preparation.

The nanodispersion of octdecyltrimethylammonium

modified MMT (qC18-MMT) in neat PLA

also promotes a higher HDT. Sinha Ray et al.

examined the HDT of neat PLA si various PLACNs

with different load conditions. As seen in Fig. 90a

there is a marked increase of HDT with an

intermediate load of 0.98 MPa, from 76 8C pentru the

neat PLA to 93 8C pentru   PLACN4. This value gradually

increases with increasing clay content, si in the

case of PLACN7 with 7 wt% of OMLS, the value

increases to 111 8C.

On the other hand, an imposed load dependence on

HDT is clearly observed in the case of PLACNs.

Fig. 90b shows the typical load dependence in

PLACN7. In the case of high load (1.81 MPa), it is

very difficult to achieve high HDT enhancement

without a strong interaction between the polymer

matrix si OMLS, as observed N6 based nanocomposites.

For PLACNs, the values of Tm do not change

significantly as compared to that of neat PLA.

Furthermore, in WAXD analyses up to 2 u¼ 708; no

large shifting or pentru mation of new peaks in the

crystallized PLACNs was observed. This suggests

that the improvement of HDT with intermediate load

originates from the better mechanical stability,

reinforcement by the dispersed clay particles, and

higher degree of crystallinity si intercalation.

The increase of HDT due to clay dispersion is a

very important property improvement pentru any polymeric

material, not only from application or industrial

point of view, but also because it is very difficult to

achieve similar HDT enhancements by chemical

modification or reinforcement by conventional filler.

4.3. Thermal stability

The thermal stability of polymeric materials is

usually studied by thermogravimetric analysis (TGA).

The weight loss due to the pentru mation of volatile

products after degradation at high temperature is

monitored as a function of temperature. When the

heating occurs under an inert gas flow, a non-oxidative

degradation occurs, while the use of air or oxygen

allows oxidative degradation of the samples. Generally,

the incorporation of clay into the polymer matrix

was found to enhance thermal stability by acting as a

Tabelul 18

HDT of PP/MMT nanocomposites si the respective unfilled PP

Organically modified mmt (wt%) HDT (8C)

PP/f-MMT PP/alkyl-MMT

3 144 ^ 5 130 ^ 7a

6 152 ^ 5 141 ^ 7b

Reproduced from Manias, Touny, Strawhecker, Lu si Chang

by permission of American Chemical Society, USA.

a C18-mmt filler, extruder processed.

b 2C18-MMT filler, twin-head mixer.

Fig. 90. (a) Organoclay (wt%) dependence of HDT of neat PLA si various PLACNs. (b) Load dependence of HDT of neat PLA si PLACN7

. Reproduced from Sinha Ray, Yamada, Okamoto si Ueda by permission of Elsevier Science Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1604

superior insulator si mass transport barrier to the

volatile products generated during decomposition.

Blumstein first reported the improved thermal

stability of a PLS nanocomposite that combined

PMMA si MMT clay. These PMMA nanocomposites

were prepared by free radical polymerization of

MMA intercalated in the clay. He showed that the

PMMA that was intercalated (d spacing increase of

0.76 nm) between the galleries of MMT clay resisted

the thermal degradation under conditions that would

otherwise completely degrade pure PMMA. TGA data

revealed that both linear si cross-linked PMMA

intercalated into MMT layers have a 40-50 8C higher

decomposition temperature. Blumstein argues that the

stability of the PMMA nanocomposite is due not only

to its different structure, but also to the restricted

thermal motion of the PMMA in the gallery.

Recently, there have been many reports concerned

with the improved thermal stability of nanocomposites

prepared with various types of OMLS and

polymer matrices . Very recently, Zanetti

et al. conducted detailed TG analyses of

nanocomposites based on EVA. The inorganic phase

was fluorohectorite (FH) or MMT, both exchanged

with octadecylammonium cation. They found that the

deacylation of EVA in nanocomposites is accelerated,

and may occur at temperatures lower than those pentru

the pure polymer or corresponding microcomposite

due to catalysis by the strongly acidic sites created by

thermal decomposition of the silicate modifier. These

sites are active when there is intimate contact between

the polymer si the silicate. Slowing down the

volatilization of the deacylated polymer in nitrogen

may occur because of the labyrinth effect of the

silicate layers in the polymer matrix

In air, the nanocomposite exhibits a significant

delay in weight loss that may derive from the barrier

effect caused by diffusion of both the volatile thermooxidation

products to the gas si oxygen from the gas

phase to the polymer. According to Gilman et al.

this barrier effect increases during volatilization

owing to the reassembly of the reticular of the silicate

on the surface.

Fig. 91 represents the TGA analysis of a phosphonium-

PS nanocomposite compared with virgin

PS. It shows that the thermal stability of the

nanocomposite is enhanced relative to that of virgin

PS , si the typical the onset temperature of

the degradation is about 50 8C higher pentru the

nanocomposites. From Fig. 91 it is clearly observed

that the degradation mechanism of phosphonium

nanocomposites is somehow different from the others;

there is a second step in the degradation. This second

step accounts pentru   about 30% of the degradation of the

phosphonium-PS nanocomposite, si must be attributed

to some interaction between the clay si the PS

that serves to stabilize the nanocomposite. The most

likely explanation is that the higher decomposition

temperature of the phosphonium clay provides pentru the

formation of char at a more opportune time to retain

the PS. In the case of ammonium clays, char

formation occurs earlier si can be broken up by

the time the polymer degrades.

The variation of the temperature at which 10%

degradation occurs pentru   all three nanocomposites is

shown as a function of the amount of clay in Fig. 92

. Even with as little as 0.1 wt% of clay present in

the nanocomposite, the onset temperature was significantly

increased.

Fig. 93 [275] shows the TGA results pentru   pure PSF

and pentru   nanocomposites containing 1 si 5 wt% of the

OMLS. The approximate decomposition temperatures

of these three materials were 494, 498 si 513 8C,

respectively. There were significant increases in

thermal stability resulting from the exfoliated clay

platelets, which may be due to kinetic effects, with the

platelets retarding diffusion of oxygen into the

polymer matrix.

The thermal stability of the PCL-based composites

has also been studied by TGA. Generally, the

degradation of PCL fits a two-step mechanism

; first random chain scission through

Fig. 91. TGA curves pentru   polystyrene, PS si the nanocomposites

. Reproduced from Zhu, Morgan, Lamelas si Wilkei by

permission of American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1605

pyrolysis of the ester groups, with the release of CO2,

H2O si hexanoic acid, then in the second step, 1 -

caprolactone (cyclic monomer) is pentru med as a result of

an unzipping depolymerization process. The thermograms

of nanocomposites prepared with Mont-Alk and

pure PCL recovered after clay extraction are presented

in Fig. 94 [171]. Both intercalated si exfoliated

nanocomposites show higher thermal stability when

compared to the pure PCL or to the corresponding

microcomposites. The nanocomposites exhibit a 25 8C

high in decomposition temperature at 50% weight loss.

The shift of the degradation temperature may be

ascribed to a decrease in oxygen si volatile

degradation products permeability/diffusivity due to

the homogeneous incorporation of clay sheets, to a

barrier of these high-aspect ratio fillers, si char

formation. The thermal stability of the nanocomposites

systematically increases with increasing clay, up to a

loading of 5 wt%.

Different behavior is observed in synthetic biodegradable

aliphatic polyester BAP/OMLS nanocomposite

systems, in which the thermal degradation

temperature si thermal degradation rate systematically

increases with increasing amounts of OMLS

up to 15 wt% . The TGA results pentru   neat BAP

and various nanocomposites are presented in Fig. 95

Fig. 94. Temperature dependence of the weight loss under an air

flow pentru   neat PCL si PCL nanocomposites containing 1, 3, 5,

and 10 wt% (relative to inorganics) of MMT-Alk

Reproduced from Lepoittevin, Devalckenaere, Pantoustier, Alexandre,

Kubies, Calberg, Jerome si Dubois by permission of

Elsevier Science Ltd, UK.

Fig. 93. TGA curves (relative weight loss as a function of

temperature) pentru   (a) pure polysulfone, (b) nanocomposite with

1 wt% clay, si (c) nanocomposite with 5 wt% clay

Reproduced from Sur, Sun, Lyu si Mark by permission of

Elsevier Science Ltd, UK.

Fig. 92. Temperature of 10% mass loss pentru nanocomposites as a

function of the fraction of clay . Reproduced from Zhu, Morgan,

Lamelas si Wilkei by permission of American Chemical Society,

USA

Fig. 95. TGA of BAP/organically modified MMT with different

organoclay . Reproduced from Lim, Hyun, Choi si Jhon by

permission of American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1606

Like PS-based nanocomposites, a small amount of

clay also increased the residual weight of BAP/

OMMT because of the restricted thermal motion of

the polymer in the silicate layers. The residual weight

of various materials at 450 8C increased in the order

BAP , BAP03 , BAP06 , BAP09 , BAP15 (here

number indicates wt% of clay). These improved

thermal properties are also observed in other systems

like SAN , the intercalated nanocomposite

prepared by emulsion polymerization.

The role of clay in the nanocomposite structure may

be the main reason pentru   the difference in TGA results of

these systems compared to the previously reported

systems. The clay acts as a heat barrier, which

enhances the overall thermal stability of the system,

as well as assist in the pentru mation of char after thermal

decomposition. In the early stages of thermal

decomposition, the clay would shift the decomposition

to higher temperature. After that, this heat barrier effect

would result in a reverse thermal stability. In other

words, the stacked silicate layers could hold accumulated

heat that could be used as a heat source to

accelerate the decomposition process, in conjunction

with the heat flow supplied by the outside heat source.

4.4. Fire retardant properties

The Cone calorimeter is one of the most effective

bench-scale methods pentru   studying the fire retardant

properties of polymeric materials. Fire-relevant properties

such as the heat release rate (HRR), heat peak

HRR, smoke production, si CO2 yield, are vital to

the evaluation of the fire safety of materials.

In 1976 Unitika Ltd, Japan, first presented the

potential flame retardant properties of N6/layered

silicate nanocomposites . Then in 1997 Gilman

et al. reported detailed investigations on flame

retardant properties of N6/layered silicate nanocomposite

. Subsequently, they chose various types of

nanocomposite materials si found similar reductions

in flammability . Recently, Gilman reviewed

the flame retardant properties of nanocomposites in

detail . Since the decreased flammability of

nanocomposites is one of the most important properties,

the results of some of the most recent studies on

flame retardant properties of nanocomposites are

reported in the following.

Tabelul 19 represents the cone calorimeter data of

three different kinds of polymer si their nanocomposites

with MMT. As shown in Tabelul 19, all of the

MMT-based nanocomposites reported here exhibit

reduced flammability. The peak HRR is reduced by

50-75% pentru   N6, PS, si PP-g-MA nanocomposites

. According to the authors, the MMT must be

nanodispersed pentru   it to affect the flammability

of the nanocomposites. However, the clay need

not be completely delaminated. In general, the

nanocomposites' flame retardant mechanism involves

Tabelul 19

Cone calorimeter data of various polymers si their nanocomposites with OMLS

Sample

(structure)

% residue yield

Peak HRR

(kW/m2)

(D%)

Mean HRR

(kW/m2)

(D%)

Mean Hc

(MJ/kg)

Mean SEA

(m2/kg)

Mean CO yield

(kg/kg)

N6 1 1010 603 27 197 0.01

N6 nanocomposite 2% (delaminated) 3 686 (32) 390 (35) 27 271 0.01

N6 nanocomposite 5% (delaminated) 6 378 (63) 304 (50) 27 296 0.02

PS 0 1120 703 29 1460 0.09

PS-silicate mix 3% (immiscible) 3 1080 715 29 1840 0.09

PS-nanocomposite 3% (intercalated/delaminated) 4 567 (48) 444 (38) 27 1730 0.08

PSw/DBDPO/Sb2O3) 30% 3 491 (56) 318 (54) 11 2580 0.14

PpgMA 5 1525 536 39 704 0.02

PpgMA-nanocomposite 2% (intercalated/delaminated) 6 450 (70) 322 (40) 44 1028 0.02

PpgMA-nanocomposite 4% (intercalated/delaminated) 12 381 (75) 275 (49) 44 968 0.02

Heat flux, 35 kW/m2. Hc; specific heat of combustion; SEA, specific extinction area. Peak HRR, mass loss rate, si SEA data,

measured at 35 kW/m2, are reproducible to within ^10%. The carbon monoxide si heat of combustion data are reproducible to within

^15%. Reproduced from Gilman, Jackson, Morgan, Harris Jr, Manias, Giannelis, Wuthemow, Hilton si Phillips by permission of

American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1607

a high-performance carbonaceous-silicate char, which

builds up on the surface during burning. This insulates

the underlying material si slows the mass loss rate of

decomposition products.

In a recent study, Zhu et al. reported the fire

retardant properties of PS/MMT nanocomposites,

which were prepared using three different types of

new organically modified MMT (see Fig. 25). They

initially used phosphonium salt pentru the modification of

clay, si then examined the differences between

organo ammonium si phosphonium salt treatments

of clay fillers in nanocomposites towards thermal

stability. The peak HRR pentru   PS si the three

nanocomposites is also shown graphically in Fig. 96

As mentioned above, the suggested mechanism by

which clay nanocomposites function involves the

formation of a char that serves as a barrier to both

mass si energy transport . As the fraction of clay

increases, the amount of char that can be pentru med

increases, si the rate at which heat is released

decreases. One of these nanocomposites, OH-16, is

mostly intercalated. This yields a slight reduction in

the rate of heat release compared with the other two

systems, which contain a significant exfoliated

fraction. This observation again supports the suggestion

that an intercalated material is more effective than

an exfoliated material in fire retardance

In contrast, the decrease in the rate of heat release

corresponds to (1) a decrease in mass loss rate si the

amount of energy released by the time PS has ceased

burning, si (2) a modest increase in the time at

which the peak heat release is reached. The production

of a char barrier must serve to retain some of

the PS, si thus both the energy released si the mass

loss rate decrease. The amount of smoke evolved and

specific extinction area also decrease with the

formation of the nanocomposites. There is some

variability in the smoke production. Although it is

observed that the pentru mation of the nanocomposites

reduces smoke production, the presence of additional

clay does not continue this smoke reduction.

4.5. Gas barrier properties

Clays are believed to increase the barrier properties

by creating a maze or 'tortuous path' (see Fig. 97) that

retards the progress of the gas molecules through the

matrix resin. The direct benefit of the pentru mation of

such a path is clearly observed in polyimide/clay

nanocomposites by dramatically improved barrier

properties, with a simultaneous decrease in the

thermal expansion coefficient . The

polyimide/layered silicate nanocomposites with a

small fraction of OMLS exhibited reduction in the

permeability of small gases, e.g. O2, H2O, He, CO2,

and ethylacetate vapors . pentru   example, at 2 wt%

clay loading, the permeability coefficient of water

vapor was decreased ten-fold with synthetic mica

relative to pristine polyimide. By comparing nanocomposites

made with layered silicates of various

aspect ratios, the permeability was seen to decrease

with increasing aspect ratio.

Oxygen gas permeability has been measured pentru

near to exfoliated PLA/synthetic mica nanocomposites

prepared by Sinha Ray et al. . The relative

permeability coefficient value, i.e. PPLACN=PPLA

where PPLACN si PPLA are the nanocomposite and

pure PLA permeability coefficient, respectively, is

plotted as a function of the wt% of OMLS in

Fig. 98. The data are analyzed with the Nielsen

theoretical expression , allowing prediction of

Fig. 96. Peak HRRs pentru   PS si the three nanocomposites

Reproduced from Zhu, Morgan, Lamelas si Wilkei by permission

of American Chemical Society, USA. Fig. 97. pentru mation of tortuous path in PLS nanocomposites.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1608

gas permeability as a function of the length si width

of the filler particles, as well as their volume fraction

in the PLA-matrix.

The H2O-vapor permeability pentru   the PUU/OMLS

nanocomposites is presented in Fig. 99 in terms of

Pc=Po; the ratio of the permeability coefficient of the

nanocomposite ðPcÞ to that of the neat PUU ðPoÞ

The nanocomposite pentru mation results in a dramatic

decrease in H2O-vapor transmission through the PUU

sheet. The solid lines in Fig. 99 are based on

the tortuosity model pentru   aspect ratios of 300 and

1000. The comparison between the experimental

values si the theoretical model suggests a gradual

change in the effective aspect ratio of the filler.

Although the enhancement in barrier properties in

nanocomposites is well known, the dependence on

factors such as the relative orientation si dispersion

(intercalated, exfoliated or some intermediate) is not

well understood. Very recently, Bharadwaj

addressed the modeling of barrier properties in

PLS nanocomposites based completely upon

the tortuosity arguments described by Nielsen

The correlation between the sheet length, concentration,

relative orientation, si state of aggregation

is expected to provide guidance in the design of

better barrier materials using the nanocomposite

approach.

The presence of filler, spherical, plate, cylindrical,

etc. introduces a tortuous path pentru a diffusing

penetrant. The reduction of permeability arises from

the longer diffusive path that the penetrants must

travel in the presence of filler, as shown in the inset in

Fig. 100. A sheet-like morphology is particularly

efficient at maximizing the path length due to the large

length-to-width ratio, when compared to other filler

shapes such as spheres or cubes. The tortuosity factor

tis defined as the ratio of the actual distance d0 that a

penetrant must travel to the shortest distance d that it

would travel in the absence of barriers. It is expressed

in terms of the length L; width W; si volume fraction

Fig. 98. Oxygen gas permeability of neat PLA si various PLACNs

as a function of OMLS content measured at 20 8C si 90% relative

humidity. The filled circles represent the experimental data.

Theoretical fits based on Nelson tortuousity model

Reproduced from Sinha Ray, Yamada, Okamoto, Ogami and

Ueda by permission of American Chemical Society, USA.

Fig. 99. Relative H2O vapor permeability pentru the PUU nanocomposites. The nanocomposite pentru mation results in a dramatic decrease in H2O

vapor transmission through the PUU membrane. The solid lines represent the theoretical value pentru   aspect ratios ¼ 300 si 1000 . Reproduced

from Xu, Manias, Snyder, Runt by permission of American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1609

of the sheets fs as

d0

d ¼ 1 þ

L

2W

fs ð5Þ

The effect of tortuosity on the permeability is

expressed as

Ps

Pp ¼

1 2 fs

t ð6Þ

where Ps si Pp represent the permeabilities of the

nanocomposite si pure polymer, respectively. In the

present case, the polygonal aluminosilicate sheets are

approximated as disks with a mean diameter L ranging

from 30 to 200 nm si a width of 1 nm.

Although the above equations were developed to

model the diffusion of small molecules in conventional

composites, they do extremely well in reproducing

experimental results pentru   the relative permeability in

PLS nanocomposites. The key assumption of this

model is that the sheets are placed in an arrangement

such that the direction of diffusion is normal to the

direction of the sheets. Clearly, this arrangement

results in the highest tortuosity, si any deviation

from the arrangement where the sheet normal lies

perpendicular to the film plane would in fact lead to

deterioration of the barrier properties. A range of

relative sheet orientations with respect to each other

and to the plane of the film is shown in Fig. 101

4.6. Ionic conductivity

Solvent-free electrolytes are of much interest

because of their charge-transport mechanism and

their possible applications in electrochemical devices.

Fig. 100. Effect of sheet orientation on the relative permeability in exfoliated PLS nanocomposites at fs ¼ 0:05 si W ¼ 1 nm. The

illustrations show the definition of the direction of preferred orientation ðnÞ of the silicate sheet normals ð pÞ with respect to the film plane.

Illustration pentru   three values of the order parameter ðSÞ-1/2, 0, si 1 are also shown . Reproduced from Bharadwaj by permission of

American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1610

Vaia et al. studied the preparation of PEO/

layered silicate nanocomposites to fine tune the ionic

conductivity of PEO. An intercalated nanocomposite

prepared by melt intercalation of PEO (40 wt%) into

Liþ-MMT (60 wt%) was shown to enhance the

stability of the ionic conductance at lower temperatures

when compared to a more conventional PEO/

LiBF4 mixture. This improvement in conductivity is

explained by the fact that PEO is not able to

crystallize when intercalated, eliminating the presence

of crystallites, which are non-conductive. The

higher conductivity at room temperature, compared to

conventional PEO/LiBF4 electrolytes with a single

ionic conductor, makes these nanocomposites promising

new electrolyte materials. The same type ionic

conductivity behavior was observed in a poly[bis

(methoxy-ethoxy) ethoxy phosphazene/Naþ-MMT

nanocomposites prepared by Hutchison et al.

4.7. Optical transparency

Although layered silicates are microns in lateral

size, they are just 1 nm thick. Thus, when single layers

are dispersed in a polymer matrix, the resulting

nanocomposite is optically clear in visible light.

Fig. 102 presents the UV/visible transmission spectra

of pure PVA si PVA/Naþ-MMT nanocomposites

with 4 si 10 wt% MMT. The spectra show that the

visible region is not affected by the presence of the

silicate layers, si retains the high transparency of

PVA. pentru   UV wavelengths, there is strong scattering

and/or absorption, resulting in very low transmission

of UV light. This behavior is not surprising, as the

typical MMT lateral sizes are 50-1000 nm.

Like PVA, various other polymers also show

optical transparency after nanocomposite preparation

with OMLS

4.8. Biodegradability of biodegradable

polymers-based nanocomposites

Another interesting si exciting aspect of nanocomposite

technology is the significant improvement

of biodegradability after nanocomposite preparation

with OMLS. Tetto et al. first reported results

on the biodegradability of nanocomposites based on

Fig. 101. Effect of incomplete exfoliation on the relative permeability. The illustrations show the effect of having one, two, si four sheet

aggregates dispersed throughout the matrix. The plot shows the relative permeability as a function of the aggregate width at several different

lengths of the sheets at fs ¼ 0:05 . Reproduced from Bharadwaj by permission of American Chemical Society, USA.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1611

PCL, reporting that the PCL/OMLS nanocomposites

showed improved biodegradability compared to pure

PCL. The improved biodegradability of PCL after

nanocomposites pentru mation may be due to a catalytic

role of the OMLS in the biodegradation mechanism,

but it this is not clear.

Recently, Lee et al. reported the biodegradation

of aliphatic polyester-based nanocomposites

under compost. Parts a si b of Fig. 103 show the

clay content dependence of the biodegradation of

APES-based nanocomposites prepared with two

different types of clays. They assumed that the

retardation of biodegradation was due to the improvement

of the barrier properties of the aliphatic APSE

after nanocomposite preparation with clay. However,

they do not provide permeability data.

Very recently, Sinha Ray et al. reported

the biodegradability of neat PLA si the corresponding

nanocomposites prepared with octadecyltrimethylammonium-

modified MMT (C18C3-MMT), along

with a detailed mechanism of the degradation. The

compost used was prepared from food waste, si tests

were carried out at a temperature of (58 ^ 2)8C.

Fig. 104a shows the recovered samples of neat PLA

and PLACN4 (C18C3-MMT ¼ 4 wt%) from compost.

The decrease in Mw si residual weight

percentage of the initial test samples are also reported

in Fig. 104b. Obviously, the biodegradability of neat

PLA is significantly enhanced after nanocomposite

preparation with C18C3-MMT. Within one month,

both the extent of Mw si the extent of weight loss are

at the same level pentru   both neat PLA si PLACN4.

However, after one month, a sharp change occurs in

the weight loss of PLACN4, si within 2 months it is

completely degraded by compost.

The presence of terminal hydroxylated edge

groups in the silicate layers may be one of the factors

responsible pentru   this behavior. In the case of PLACN4,

the stacked (,4 layers) si intercalated silicate layers

are homogeneously dispersed in the PLA matrix (from

TEM image), si these hydroxy groups start heterogeneous

hydrolysis of the PLA matrix after absorbing

water from the compost. This process takes some time

to start. pentru   this reason, the weight loss si degree of

hydrolysis pentru   PLA si PLACN4 are almost the same

up to 1 month (see Fig. 104b). However, after 1 month

there is a sharp weight loss in the case of PLACN4

compared to that of PLA. That means that 1 month is

the critical timescale to start heterogeneous hydrolysis,

and due to this type of hydrolysis, the matrix

decomposes into very small fragments si eventually

disappears with the compost. This assumption was

confirmed by conducting the same experimental

procedure with PLACN prepared with dimethyldioctdecylammonium

salt modified synthetic mica, which

Fig. 102. UV-vis transmittance spectra of PVA si PVA/MMT

nanocomposites containing 4 si 10 wt% MMT . Reproduced

from Strawhecker si Manias by permission of American Chemical

Society, USA

Fig. 103. Biodegradability of APES nanocomposites with: (a) Closite 30B si (b) Closite 10A . Reproduced from Lee, Park, Lim, Kang,

Li, Cho si Ha by permission of Elsevier Science Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1612

has no terminal hydroxylated edge group. The same

degradation tendency was found with PLA

A respirometric test has also been used to study the

degradation of the PLA matrix in a compost environment

at (58 ^ 2)8C . pentru   this test compost

was made from bean-curd refuse, food waste, and

cattle feces. Unlike the weight loss, which reflects the

structural changes in the test sample, CO2 evolution

provides an indicator of the ultimate biodegradability

of PLA in PLACN4 (prepared with (N(coco alkyl)N,N-

[bis(2-hydroxyethyl)]-N-methylammonium modified

synthetic mica), via mineralization, of the samples.

Fig. 105 shows the time dependence of the degree of

biodegradation of neat PLA si PLACN4, indicating

that the biodegradability of PLA in PLACN4 is

enhanced significantly. The presence of MEE clay

may cause a different mode of attack on the PLA

component due to the presence of hydroxy groups.

Details regarding the mechanism of biodegradability

can be found in relevant literature

4.9. Other properties

PLS nanocomposites also show improvement in

most general polymeric properties. pentru example, in

addition to the decreased permeability of liquids and

gases, nanocomposites also show significant improvement

in solvent uptake. Scratch resistance is another

property that is strongly enhanced by the incorporation

of layered silicates

The potential to use PANI-based nanocomposites

as electrorheologically sensitive fluids , or to use

Fig. 104. (a) Real picture of biodegradability of neat PLA si PLACN4 recovered from compost with time. Initial shape of the crystallized

samples was 3 £ 10 £ 0.1 cm3. (b) Time dependence of residual weight, Rw si of matrix, Mw of PLA si PLACN4 under compost at

(58 ^ 2) 8C . Reproduced from Sinha Ray, Yamada, Okamoto si Ueda by permission of Elsevier Science Ltd, UK.

S. Sinha Ray, M. Okamoto / Prog. Polym. Sci. 28 (2003) 1539-1641 1613

the combination of dispersed layered silicates in a

liquid crystal medium is also an attractive application.

This could result in a stabelul electro-optical device that

is capable of exhibiting a bistabelul si reversible

electro-optical effect between an opaque si transparent

state

Finally, nanocomposites have been used in highly

technical areas such as in the improvement of ablative

properties in aeronautics


Document Info


Accesari: 11124
Apreciat: hand-up

Comenteaza documentul:

Nu esti inregistrat
Trebuie sa fii utilizator inregistrat pentru a putea comenta


Creaza cont nou

A fost util?

Daca documentul a fost util si crezi ca merita
sa adaugi un link catre el la tine in site


in pagina web a site-ului tau.




eCoduri.com - coduri postale, contabile, CAEN sau bancare

Politica de confidentialitate | Termenii si conditii de utilizare




Copyright © Contact (SCRIGROUP Int. 2024 )