influência da técnica e material restaurador no comportamento biomecânico...

47
Universidade Estadual de Campinas Faculdade de Odontologia de Piracicaba RODRIGO BARROS ESTEVES LINS Influência da técnica e material restaurador no comportamento biomecânico em Restaurações Classe II Influence of the restorative technique and material on the biomechanical behavior of Class II restorations Piracicaba 2017

Upload: others

Post on 28-Jan-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

  • Universidade Estadual de Campinas

    Faculdade de Odontologia de Piracicaba

    RODRIGO BARROS ESTEVES LINS

    Influência da técnica e material restaurador no comportamento

    biomecânico em Restaurações Classe II

    Influence of the restorative technique and material on the biomechanical

    behavior of Class II restorations

    Piracicaba

    2017

  • RODRIGO BARROS ESTEVES LINS

    Influência da técnica e material restaurador no comportamento

    biomecânico em Restaurações Classe II

    Influence of the restorative technique and material on the biomechanical

    behavior of Class II restorations

    Dissertação apresentada à Faculdade de

    Odontologia de Piracicaba da Universidade

    Estadual de Campinas como parte dos requisitos

    exigidos para a obtenção do título de Mestre em

    Clínica Odontológica, Área de Dentística.

    Dissertation presented to the Piracicaba Dental

    School of the University of Campinas in partial

    fulfillment of the requirements for the degree of

    Master in Odontological Clinic, in Restorative

    Dentistry area.

    Orientador: Professor Doutor Luis Roberto Marcondes Martins

    ESTE EXEMPLAR CORRESPONDE À VERSÃO FINAL

    DA DISSERTAÇÃO DEFENDIDA PELO ALUNO

    RODRIGO BARROS ESTEVES LINS E ORIENTADA

    PELO PROF. DR. LUIS ROBERTO MARCONDES

    MARTINS.

    Piracicaba

    2017

  • Dedicatória

    À Deus e à Nossa Senhora,

    Além de dedicar a minha dissertação à Deus e à Nossa Senhora, dedico todo o meu

    mestrado, meus estudos, minha vida profissional e pessoal.

    Aos meus pais,

    Dedico este trabalho de obtenção do título de Mestre aos meus maiores Mestres: ao

    meu pai (José Lins) e à minha mãe (Mariângela Lins).

    Por terem me apoiado em todas as minhas decisões, incluindo sair de casa para

    correr atrás de um futuro profissional que ainda almejo em conquistar. Vocês são meus maiores

    exemplos para seguir em busca dos meus ideais, segundo à vontade de Deus.

    Muito obrigado por todo apoio e carinho. Amo vocês!

  • Agradecimentos

    Aos meus familiares,

    Agradeço a todos os meus familiares pelo apoio e incentivo dados a mim nesta nova

    fase de estudos e aprendizados, principalmente às minhas avós.

    Em especial aos meus irmãos José Renato Lins e Rafael Lins. A saudade é imensa,

    mas sei que vocês estão e estarão sempre torcendo por mim.

    A minha querida sobrinha Isabelle Lins, que sinto tanta falta. Estar longe e não

    poder acompanhar pessoalmente seu crescimento é extremamente penoso, mas ter o seu carinho

    em todas as viagens de retorno à João Pessoa é mais do que reconfortante.

    Aos meus amigos Pessoenses,

    Aos tantos amigos que fiz durante a vida, incluindo amigos da época de escola

    (Augusto Ygor, Paulo Victor e Suzane Souza) e universidade (Alana Dantas, Amanda Solano,

    Evllon Sá, Larissa Fernandes, Maíra Ramalho, Mariana Figueiredo, Raissa Marçal, Priscilla

    Leite e tantos outros) os quais mantemos contato por redes sociais, agradeço por fazerem parte

    da minha vida.

    Ao meu orientador Prof. Dr. Luís Roberto Marcondes Martins,

    Obrigado por ter me escolhido na seleção de mestrado e por ter acreditado no meu

    potencial para desenvolver trabalhos científicos juntos. Nestes dois anos foram tempos de

    muitos aprendizados que servirão para a minha vida. Tê-lo como orientador é mais do que um

    exemplo.

    Aos Professores da Dentística,

    A todos os professores da Dentística-FOP, por todos os ensinamentos e atenção

    prestados nestes dois anos de convivência diária. Grandes mestres que nos dão muito orgulho

    de suas competências profissionais.

  • Aos funcionários da FOP-UNICAMP

    Não poderia deixar de agradecer aos funcionários da FOP-UNICAMP que nos

    prestam tantos auxílios nas mais diversas funções, desde aos funcionários da secretaria às

    senhoras da limpeza.

    Aos amigos Fopianos/Pessoenses,

    Aos amigos Pessoenses que estudam na FOP-UNICAMP (Bruno Mariz, Emerson

    Tavares, Jaiza Araújo, Jossaria Sousa, Laíse Lima, Marina Moreno, Mayara Abreu, Renally

    Wanderley) que tornam a FOP um pouco nordestina e que fazem os nossos dias de estudo mais

    divertidos.

    Aos amigos da FOP-UNICAMP

    Aos grandes amigos que conquistei nestes dois anos de mestrado (Amanda

    Bandeira, Camilla Fraga, Carolina Rangel, Filipe Martins, Heloisa Pantaroto, Jairo Cordeiro,

    Manoelito Silva, Mariana Barbosa, Olívia Figueiredo, Rafaela Videira, Rahyza Freire, Renato

    Machado, Thiago Bessa). Agradeço por todo o companheirismo e apoio nos momentos de

    felicidade e principalmente naqueles momentos de preocupação ou estresse inevitáveis à esta

    fase.

    Aos amigos da Dentística/FOP-UNICAMP

    Aos amigos da Dentística que dividimos diariamente conhecimentos científicos

    (Renata Pereira, Bruna Guerra, Caroline Mathias, Mayara Zaghi e Marina Faria) e em especial

    a Josué Pierote, Mariana Flor e Maicon Sebold pela grande amizade.

    À Cristiane Yanikian

    Um agradecimento especial à Cristiane Yanikian. Ela que me pôs para trabalhar no

    laboratório no meu primeiro dia de mestrado, mas que soube como ninguém me ensinar e

    orientar por todos os obstáculos enfrentados durante o mestrado. É sem dúvida, um exemplo

    que pretendo seguir.

  • À Elis Lira e Louise Dornelas

    À minha família Piracicabana; às minhas colegas de graduação e pós-graduação; às

    minhas grandes amigas: Elis Lira e Louise Dornelas. Mais de 7 anos de amizade. Não sei

    expressar o quanto vocês foram e são importantes para mim. Ter ingressado na pós-graduação

    com vocês foi, certamente, o ‘combustível’ necessário para suprir a saudade de casa. Tenho

    muito orgulho pelas profissionais que são e pelas conquistas que não cessam em acontecer.

    Ao Professor Doutor João Carlos Ramos e à Professora Doutora Alexandra Vinagre,

    Agradeço aos Professores da Faculdade de Medicina Dentária da Universidade de

    Coimbra-Portugal, que me acolheram durante o meu intercâmbio na reta final do meu mestrado.

    Exemplo de dedicação à odontologia, orientadores e mestres incríveis. Profissionais que

    durante 4 meses me deram suporte e me concederam grandes conhecimentos. Obrigado por

    todo o tempo e atenção depositados à mim.

    À Universidade Estadual de Campinas

    À Faculdade de Odontologia de Piracicaba

    Ao Programa de Clínica Odontológica

    À CAPES

  • RESUMO

    O advento de materiais restauradores de baixa contração vem ganhando espaço no mercado e

    despertando interesse de cirurgiões-dentistas devido as suas características técnicas, economia

    de tempo clínico e por ser um material mais estável dentro da cavidade oral. O objetivo deste

    estudo foi avaliar o comportamento biomecânico de restaurações extensas classe II mésio-

    ocluso-distal (MOD) com técnica direta (compósitos convencional e de baixa contração)

    avaliando a tensão de contração de polimerização, deflexão de cúspide, resistência à fratura e o

    padrão de fratura. Sessenta terceiros molares humanos hígidos foram selecionados e

    randomizados em quatro grupos: resina composta Z100 (Z100); Tetric N-Ceram Bulk Fill

    (TNC); Filtek Bulk Fill (FBF); Aura Ultra Universal Restorative Material (ABF). Cavidades

    Classe II MOD foram confeccionadas com dimensões padronizadas por meio de uma máquina

    de preparo cavitário. As resinas compostas bulk fill foram inseridas na cavidade em um único

    incremento de 4 mm e a resina composta convencional em três incrementos de forma oblíqua e

    fotopolimerizadas por um LED de alta potência. Os sensores FBG foram inseridos na interface

    adesiva a fim de avaliar a tensão de contração do material resinoso durante a fotopolimerização

    (n=5). Além disso, extensômetros foram fixados nas bases das cúspides para medirem a

    deformação em três momentos (n=10): durante o procedimento restaurador, sob carregamento

    de compressão axial em máquina de ensaio universal até 100N e até gerar a fratura do conjunto

    dente-restauração. O tipo de fratura foi classificado conforme o seu padrão analisado em

    Microscopia Eletrônica de Varredura (MEV): I-fratura em resina composta; II-fratura em resina

    composta e estrutura dental coronal; III-fratura em resina composta e estrutura dental cervical

    com possibilidade de reparo periodontal; IV-fratura radicular sem reparo. Foi realizado teste de

    homogeneidade dos dados (Shapiro-Wilk, p>0.05) para todos os grupos, seguido do teste

    ANOVA um fator com teste post-hoc de Tukey para a avaliação dos sensores FBG, deformação

    de cúspide e resistência à fratura. O padrão de fratura foi analisado pelo teste qui-quadrado. A

    avaliação da tensão de contração pelos sensores de Bragg mostrou maiores médias para o grupo

    Z100 diferindo estatisticamente de todos os outros grupos (p

  • fratura do TNC foi maior que os demais, porém, estatisticamente significante apenas com o

    grupo Z100 (p

  • ABSTRACT

    The appearance of low contraction restorative materials has been introduced and increasing

    attention from dental surgeons due to technical features, such as chair time economy and stable

    within the oral cavity. The purpose of this study was to evaluate the biomechanical behavior of

    extensive Class II mesial-occlusal-distal (MOD) restorations with direct technique

    (conventional and low-shrinkage resin composites) by using fiber Bragg grating sensors (FBG)

    and extensometry assessing the polymerization contraction stress, cuspal deformation, fracture

    resistance and fracture pattern. Sixty human caries-free third molars were selected and

    distributed randomly into six groups: Z100 restorative material (Z100); Tetric N-Ceram Bulk

    Fill composite (TNC); Filtek Bulk Fill composite (FBF); Aura Ultra Universal Restorative

    Material (ABF). Class II cavities MOD was standardized in an abrasion standardizing

    equipment. The bulk fill materials were inserted in bulk increment of 4 mm and the

    conventional resin composite in three oblique ones by poly-wave LED light-curing unit. The

    optical FBG sensors were fixed at adhesive interface due to evaluated the shrinkage stress of

    resinous material during the fotopolymerization (n=5). Moreover, strain gauges were fixed on

    cuspal base to measure the deformation at three times (n=10): during restorative procedure,

    subjected to axial compressive in universal testing machine up to 100N and to occur the sample

    fracture. The failure mode was rated according to their standard analyzed in Scanning Electron

    Microscopy (SEM): I-fracture at resin composite; II-fracture at resin composite and coronal

    tooth structure; III-fracture at resin composite and cervical tooth structure with possible

    periodontal repair; IV-root fracture beyond repair. The statistical analysis was performed for

    homogeneity distribution (Shapiro-Wilk, p>0.05), followed by parametric statistical tests and

    one-way ANOVA with post-hoc Tukey’s test. Data on fracture mode were submitted to Chi-

    square test. The optical FBG sensors evaluation showed that Z100 presented the highest means

    of stress shrinkage (p

  • Keywords: Composite Resins. Fiber Optical Technology. Polymerization.

  • SUMÁRIO

    1 INTRODUÇÃO 14

    2 ARTIGO: Biomechanical Behavior of Class II Restoration Using Fiber

    Bragg Grating Sensors and Extensometry

    17

    3 CONCLUSÃO 36

    REFERÊNCIAS 37

    APÊNDICE 1: Metodologia Ilustrada 42

    ANEXO 1: Certificado - Comitê de Ética em Pesquisa 46

    ANEXO 2: Comprovante de submissão do artigo 47

  • 14

    1 INTRODUÇÃO

    A exigência por parte da população em adquirir estética do sorriso e saúde bucal

    satisfatória tem influenciado no advento de materiais cada vez mais estéticos na clínica

    odontológica. Somado a isso, o desenvolvimento tecnológico impulsiona a geração de novos

    materiais restauradores e técnicas adesivas mais eficazes e duradoras (Pashley et al., 2011) com

    propriedades físicas e ópticas semelhantes aos dos dentes naturais (Nahsan et al., 2012). A partir

    desta constante inovação, podemos caracterizar o estado da arte dos materiais odontológicos

    (Ferracane, 2011).

    A tendência atual de realizar cada vez mais procedimentos preventivos na saúde

    pública para impulsionar a diminuição na prevalência da doença cárie não garante ou não

    impede a necessidade de substituição de restaurações insatisfatórias, que a partir deste ciclo

    restaurador acarreta em perda da estrutura dental. Com isso, na clínica odontológica são

    frequentemente realizadas restaurações em cavidades classe II MOD amplas (Deliperi e

    Bardwell, 2002; Bonecker et al., 2013).

    A resina composta, ao longo dos anos, ganhou grande destaque e aplicabilidade na

    clínica odontológica para diversos procedimentos (Ferracane, 2011). Esta é comumente

    utilizada em restaurações para dentes anteriores e posteriores, com excelente desempenho

    clínico, comprovado em estudos longitudinais (Demarco et al., 2012; Cetin et al., 2013; Ozakar-

    Ilday et al., 2013; Demarco et al., 2015; Van Dijken e Pallesen, 2016). Este sucesso clínico,

    principalmente em restaurações diretas com compósitos microhíbridos (Da Rosa Rodolpho et

    al., 2011; Demarco et al., 2015) estão relacionados às propriedades mecânicas favoráveis, como

    adequada resistência à fratura (Ferracane, 2011). Entretanto, algumas desvantagens também são

    observadas, como: a necessidade de inserção do material de forma incremental, a qual demanda

    maior tempo clínico e incorporação de bolhas no corpo da restauração; alta contração de

    polimerização e, consequente, geração de tensão, provocando deslocamento do compósito

    resinoso a partir da interface adesiva, criando um ‘gap’ marginal, e portanto, sensibilidade pós-

    operatória, microinfiltração, cárie secundária, pigmentação marginal e falha na restauração

    (Casselli et al., 2013; Campos et al. 2014; Dominguez et al., 2014; Benetti et al., 2015; Fronza

    et al., 2015).

    A contração de polimerização ocorre ao longo do processo de conversão dos

    monômeros em polímeros, através de forças covalentes fortes, promovendo diminuição da

    distância molecular. Além disso, a tensão ocorre na transição da fase gel para um estado amorfo

  • 15

    isotrópico quando fotoativado (Lee et al., 2013), contribuindo para a formação de tensões ao

    longo da interface adesiva das restaurações (Loguercio et al., 2004).

    Algumas alterações nas resinas compostas têm sido propostas com o objetivo de

    diminuir os efeitos indesejados da forte tensão gerada na interface dente/restauração, como a

    composição química ou estrutural do monômero da matriz resinosa; a quantidade, forma e/ou

    tratamento superficial da partícula de carga, além de modificação no sistema iniciador (El-

    Damanhoury e Platt, 2014; Kwon et al., 2015; Orlowski et al., 2015; Atalay et al., 2016).

    Recentemente, um grupo de resinas compostas de baixa contração foram

    desenvolvidas e denominadas “bulk-fill”. Estas permitem a inserção de único incremento de até

    4-5 mm de material (Czasch e Ilie, 2013); possuem a capacidade de gerar menor tensão de

    contração de polimerização; reduzem a deflexão de cúspide (Moorthy et al., 2012); apresentam

    menor tempo e alta reatividade à fotopolimerização, devido ao aumento da translucidez e,

    portanto, permitem uma maior penetração de luz e profundidade de polimerização (Roggendorf

    et al, 2011; Leprince et al, 2014; Orlowski et al., 2015); além disso, apresentam a vantagem em

    diminuir o tempo de execução clínico (Campos et al, 2014).

    Na literatura científica encontram-se diversos testes científicos destinados a

    avaliarem o comportamento de materiais resinosos, como por exemplo: análise por elementos

    finitos (Anatavara; Sitthiseripratip; Senawongse, 2016), tensômetro conectado a um sistema de

    cantilever com sensores (Kalliecharan et al., 2016), máquinas de ensaio universal (Witzel et al.,

    2005), sistema ‘ring slitting’ (Park e Ferracane, 2005; Park e Ferracane, 2006), sensores de rede

    de Bragg em fibra óptica (Rajan et al., 2016; Vinagre et al., 2016; Umesh et al., 2016),

    extensometria (Bicalho et al., 2014; Pereira et al., 2015), dentre outros.

    O sensor de rede de Bragg em fibra óptica possui a característica de ser

    fotossensível e a capacidade de verificar variações, como: temperatura, contração, expansão e

    pressão a partir da alteração do comprimento de onda (Yeh et al., 2011; Umesh et al., 2016).

    Esta tecnologia tem-se tornado uma das técnicas mais usadas em sistemas de avaliação

    biomecânica e reabilitação em diversas áreas científicas, como: automotiva, aeronáutica,

    medicina, odontologia e especificamente a biomecânica dental, incluindo estudos in vitro e in

    vivo (Al-Fakih; Abu Osman; Mahamd Adikan, 2012; Umesh et al., 2016).

    Um método adicional para compreender o comportamento biomecânico de um

    material resinoso a partir da alteração da deflexão de cúspides é por meio da extensometria, a

    qual é um teste mecânico não destrutivo que avalia diretamente as características do processo

  • 16

    restaurador, ou seja, detecta alterações do comportamento da estrutura dental resultantes das

    forças atuantes na camada adesiva, que podem ser influenciadas pelo material restaurador,

    sistema adesivo, técnica restauradora a ser adotada e o tamanho da cavidade (Bicalho et al.,

    2014; Pereira et al., 2016).

    Encontram-se na literatura científica estudos sobre resinas compostas ‘bulk fill’,

    contudo, o conhecimento sobre o comportamento biomecânico na interface adesiva das mesmas

    ainda se apresenta escarça. Pensando nisso, o objetivo do presente estudo foi avaliar o

    comportamento biomecânico de restaurações extensas classe II mésio-ocluso-distal (MOD)

    com técnica direta (compósitos convencional e de baixa contração) avaliando a tensão de

    contração de polimerização, deflexão de cúspide, resistência à fratura e o padrão de fratura.

  • 17

    2 ARTIGO:

    Biomechanical Behavior of Class II Restoration Using Fiber Bragg Grating Sensors and

    Extensometry

    Rodrigo Barros Esteves Linsa,*, Cristiane Rumi Fujiwara Yanikiana, Thiago Henrique

    Scarabello Stapea, Aline Arêdes Bicalhob, Carlos José Soaresb, Luís Roberto Marcondes

    Martinsa

    a Department of Restorative Dentistry, Piracicaba Dental School, State University of Campinas,

    Av. Limeira, 901, Piracicaba 13414-903, SP, Brazil

    b Department of Operative Dentistry and Dental Materials, Dental School, UFU – Federal

    University of Uberlândia, Av. Pará, 1720, Campus Umuarama, Uberlândia 38400-902, MG,

    Brazil

    *Corresponding author. Tel.: +55 83 988527948

    E-mail addresses: [email protected] (R.B.E.Lins), [email protected]

    (R.R.F.Yanikian), [email protected] (T.H.S.Stape), [email protected]

    (A.A.Bicalho), [email protected] (C.J.Soares), [email protected] (L.R.M.Martins)

    Funding sources

    The CAPES Foundation of Department of Education, Brazil, supported this work.

    ABSTRACT

    Objective. The purpose of this study was to assess the biomechanical behavior of extensive

    Class II mesial-occlusal-distal (MOD) restorations performed direct technique (conventional

    and low-shrinkage resin composites) by using fiber Bragg grating sensors (FBG) and research

    the polymerization contraction stress as well as evaluate the low-shrinkage resin composite

    regarding cuspal deformation and fracture resistance and pattern.

    Methods. Sixty human caries-free third molars were selected and distributed randomly into four

    groups: Z100 restorative material (Z100); Tetric N-Ceram Bulk Fill composite (TNC); Filtek

    Bulk Fill composite (FBF) and Aura Ultra Universal restorative material (ABF). In all samples

    were standardized class II MOD cavities. The bulk fill materials were inserted in bulk increment

  • 18

    and the conventional resin composite in three ones. Shrinkage stress was evaluated with optical

    FBG sensors (n=5). The cuspal deformation with extensometer was measured during

    restoration, compressive axial force and until the fracture on universal machine (n=10). The

    fracture pattern was analyzed with Scanning Electron Microscopy (SEM). The statistical

    analysis was performed for homogeneity distribution (Shapiro-Wilk, p>0.05), followed by

    parametric statistical tests and one-way ANOVA with post-hoc Tukey’s test. Data on fracture

    mode were submitted to Chi-square test.

    Results. The optical FBG sensors evaluation showed that Z100 presented the highest means of

    stress shrinkage (p

  • 19

    structure. This results in extensive Class II MOD cavity restorations, which are regularly

    performed in clinical odontology [7,8].

    The clinical success and longevity of direct restorations in posterior teeth by using

    micro-hybrid composite resins are related to favorable mechanical properties [2,9], such as

    adequate fracture resistance [10]. However, some drawbacks are found: incremental insertion

    technique, which demands longer clinic procedures; contraction due to polymerization and

    subsequent stress, with risk of dislodging of the resinous composite from the adhesive interface,

    causing marginal gap and post-procedure sensitivity; micro-infiltration; secondary caries;

    marginal pigmentation; and restoration failure [11-13]. This contraction might occur throughout

    the process of polymerization as a result of strong covalent forces, causing abridgement of

    molecular distances and tension in the passage from gel to isotropic amorphous state [14], thus

    contributing to the formation of stress at the adhesive interface of the restorations [15].

    In order to diminish the undesired effects of the composites, such as the tension

    created on the tooth/restoration interface, some chemical and structural changes in the

    composite resin composition have been proposed. This include changes in the resinous matrix,

    quantity, and shape and/or surface treatment of the inorganic particle [1]. Another group of low-

    shrinkage composite resins have been recently developed, the bulk-fill resins. They allow the

    insertion of a single increment with 4-5-mm [16] due to their capacity of generating less

    contraction stress and high reactivity to photopolymerization as result of increased

    translucency, improving light penetration and depth of polymerization [17-19]. The incremental

    restoration technique, which uses micro-hybrid conventional composite resin inserted into the

    cavity in small increments [20], is still the most adequate photopolymerization to modify the

    contraction pattern in direct restorations [21]. However, this technique has disadvantages such

    as lodging of air bubbles during incremental insertion and longer chair time [22], whereas bulk-

    fill composite resins have the advantages of controlling polymerization shrinkage stress [23],

    shortening chair time [22] and reducing cuspal deflection [24].

    The scientific literature is comprehensive about the study of bulk fill resin

    composite, however the biomechanical behavior knowledge in adhesive interface is scarce,

    wherefore, the purpose of this study was to evaluate the biomechanical behavior of extensive

    Class II mesial-occlusal-distal (MOD) restorations with direct technique (conventional and

    low-shrinkage resin composites) by using fiber Bragg grating sensors (FBG) and extensometry

    forma the polymerization contraction stress, cuspal deformation, fracture resistance and

    fracture pattern. The hypothesis tested here was: bulk-fill composite resins generate less

  • 20

    polymerization contraction stress and reduce cuspal deformation compared to conventional

    composite resins.

    2 MATERIALS & METHODS

    Four commercial resin-based composites were investigated: one conventional

    incrementally layered composite (Z100 restorative material) (positive control); three bulk-fill

    composites (Tetric N-Ceram Bulk Fill, Filtek Bulk Fill and Aura Ultra Universal restorative

    materials). Product specifications are listed in Table 1.

    All the restorative materials were light-cured by using a poly-wave LED light-

    curing unit (VALO, Ultradent Products Inc., South Jordan, UT, USA) operating with

    wavelength of 350 and 550 nm and irradiance at 995 ± 2mW/cm2 according to the

    manufacturer’s recommendations as follows: 40 seconds for each layer of conventional

    composite (three oblique increments), 20 seconds for one bulk layer of bulk-fill composites (10

    seconds for mesio-occlusal and 10 for distal-occlusal).

    2.1 Specimen Preparation

    Sixty human caries-free third molars were extracted and stored in buffered aqueous

    solution of 0.2% sodium azide at 4 ⁰C for up to 6 months. The research project was previously

    approved by the Research Ethical Committee of Piracicaba Dental School, State University of

    Campinas, Piracicaba, Brazil (protocol 127/2014).

    The teeth were cleaned by using periodontal curettes to remove organic and

    inorganic remnants, including pumice paste, water and Robson brush at low rotation before

    storage in distilled water at 4°C. The dimensions of the teeth were determined by using a digital

    calliper (Mitutoyo, Tokyo, Japan) to measure the mesio-distal (MD) and lingual-buccal (LB)

    dimensions of the occlusal surface. The occlusal surface area was determined by considering

    the molar as a rectangle with the sides determined by the MD and LB dimensions. Teeth

    selected presented occlusal surface area varying no more than 10% of the sample average.

    2.1.1 Inclusion and Simulation of Periodontal Ligament

  • 21

    The teeth were inserted into polystyrene resin to simulate the periodontal ligament.

    A 0.5 mm graphite pencil was used to mark the insertion site at 2-mm below the amelocemental

    junction, thus delineating a radicular portion to be coated with #7 molten paraffin wax. The

    teeth were fixed with sticky wax from the crown to the stem of a prosthetic liner. A radiographic

    film with central perforation compatible with the dimension of each tooth was positioned at the

    same level of the sticky wax. The mobile table of the prosthetic liner was adjusted

    perpendicularly to the long axis and on the top of the teeth. Next, a PVC cylinder with a 10-

    mm wide central perforation was attached to the radiographic film with sticky wax. The set was

    taken to a wooden plate with individual perforations for each tooth. The self-curing polystyrene

    resin was prepared and poured into the interior of the PVC cylinder. After polymerization, the

    set was removed from the rack.

    The teeth were removed from the artificial alveoli and cleaned with jets of baking

    soda and water, and the resin cylinders were polished by using a universal sanding machine

    (AROPOL-E Arotec, Cotia/São Paulo, Brazil) to eliminate the excesses. The polyether-based

    casting material (Polyether Impregum Soft 3M ESPE, St. Paul, USA) was prepared and inserted

    into the spaces corresponding to the periodontal ligament [25]. The teeth were inserted by finger

    pressure until the 2-mm mark of the amelocemental limit was aligned to the surface of the

    cylinder. After polymerization, the excesses were removed by using a scalpel blade and the

    specimens were stored in distilled water at 4°C.

    The teeth, with their respective artificial alveoli, were randomly divided into four

    groups (n = 5 to FBG sensors and n = 10 to cuspal deformation) according to material used for

    restoration.

    2.1.2 Cavity Preparation

    Cavity preparation was performed by using abrasion equipment with three

    coordinate axes millimetrically controlled (Mitutoyo Am. Corp., Ontario, Canada). Diamond

    drills 2131 and 3131 (KD Sorensen, São Paulo, Brazil) were used to prepare the cavities by

    positioning them perpendicularly to the long axis of the tooth, determining 6 degrees of

    expulsion in the surrounding and axial walls and inner round angles.

    Class II MOD cavities were standardized as follows: lingual-buccal isthmus width

    of 4 mm and occlusal depth of 3 mm, occlusal depth of 4 mm in the proximal box and 1-2 mm

    above the cemento-enamel junction of the proximal box, ending mesially and distally [26].

  • 22

    2.1.3 Insertion of Resin Composite

    The resin composite insertion of Z100 group was incremental way from oblique

    technique and thickness up until 2 mm. Three incremental were inserted, the first one was to

    buccal wall, the second to opposite wall and the last one was to reproduce the teeth occlusal

    anatomic.

    Already for TNC, FBF and ABF groups were restored with one single increment of

    4 mm thickness to fill all bulk MOD cavity.

    2.2 Fiber Bragg Grating Sensors (FBG)

    Optical FBG sensors were used to assess the shrinkage stress of resinous material

    during the photopolymerization with values expressed in micro-strain (µɛ) (n = 5). Fiber Bragg

    gratings consist of longitudinal modulations in the refractive index of optical fiber’s core region

    with a periodicity in the order of hundreds of nanometers (Smart Fibers Ltd. UK). The refractive

    index modulation allows the fundamental core mode to be coupled to a counter propagating

    core mode. Coupling occurs at a specific wavelength, λB, given by Eq. (1) – where neff is the

    effective refractive index of the fundamental core mode and Λ is the grating pitch.

    𝜆𝐵 = 2 𝑛𝑒𝑓𝑓 Λ (1)

    As the core mode is coupled to a counter propagating one by the Bragg grating, the

    optical response of a FBG can be measured in reflection. Thus, if light from a broadband light

    source is launched in a fiber in which a FBG was inscribed, a peak at λB (Bragg peak) can be

    measured in the reflection spectrum. Figure 1a shows a typical spectrum from a Bragg grating

    used in the experiments reported herein (Λ = 535.6 nm). The sensors were monitored with an

    optical spectrum interrogation analyzer (W3-2050 FBG, Smart Fibers Ltd., UK), which can

    measure tension up to 42,000 µɛ at maximum measurement frequency of 50Hz [27].

    The application of strain to a FBG causes its spectral peak to shift. This shifting is

    due to the fact that strain induces changes in the effective refractive index of the fundamental

    core mode (via strain-optic effect), neff, and variations in the grating period, Λ. By observing

    Eq. (1), it is possible to conclude that variations in neff and Λ imply in a different value for λB.

    In order to characterize the Bragg wavelength shift as a function of the applied

    strain, one have fixed the optical fiber on motorized translation stages and applied strain

  • 23

    increments to the same. Figure 1b shows the evolution of a FBG spectrum as strain increments

    are applied to the fiber. Figure 1c presents the wavelength shift as a function of the strain level.

    Measured data allowed calculating a strain sensitivity of (1.124 ± 0.005) pm/µε.

    The evaluation of resin shrinking by the employment of fiber Bragg gratings was

    carried out as follows: initially, the fiber was glued on supports and, in sequence, it was

    tensioned. As a second step, the tooth under test was elevated up to fiber level. At this point,

    the fiber could be attached on the side buccal wall of the MOD cavity at the level of the dentine

    enamel junction (Adper Single Bond 2 adhesive was employed for fixation) [28]. The

    restorative technique to be probed was then performed inside the cavity and the Bragg peak was

    followed during the curing procedure. Temperatures effects of the light and curing exotherm

    were not eliminated.

    2.3 Cuspal Deformation

    Strain gauges (PA-06-060CC-350-LEN, Excel Sensores, SP, Brazil) were used to

    measure cuspal deformation (µɛ) and they were attached parallel to the long axis of the teeth at

    the gingival wall level of the proximal box and placed on the external surface of the buccal and

    lingual cusps (n = 10). The strain gauges had an internal electrical resistance of 350Ω and gauge

    factor of 2.17. Moreover, two strain gauges were positioned to another unimpaired tooth to

    compensate for dimensional alteration due to temperature effects on half bridge scheme. The

    fixation of strain gauges were performed with 37% acid etching for 30 seconds, cyanoacrylate-

    based adhesive (Super Bonder, Loctite, Itapeví, Brazil) and digital pressure for 60 seconds and

    the wires were connected to a data acquisition device (ADS0500IP, Lynx, SP, Brazil). The

    cuspal deformation values were recorded at 4 Hz during the restorative technique and continued

    for 10 minutes after curing the last increment until stabilization [29].

    2.4 Fracture Resistance Test

    After measurement of cuspal deformation during restoration, the teeth were placed

    on universal testing machine (EMIC, 500DL, São José dos Pinhais, Brazil) and submitted to

    axial compressive loading with a metal sphere of 8 mm in diameter at speed of 0.5 mm/min up

    to 100N in order to cause fracture and assess the cuspal deformation in these moments. The

    load required (N) to cause catastrophic fracture of the specimens was recorded by a 500-N load

  • 24

    cell hardwired to a computer with control and data acquisition software (TESC 3.04, EMIC)

    [30].

    2.5 Scanning Electron Microscopy (SEM)

    The failure mode of each specimen was analyzed under the SEM (JEOL-JSM

    5600LV, Tokyo, Japan). Each sample was gold-sputter coated, submitted to analysis at 15kV

    to observe surface characteristics and classified into one of the four categories as follows: (I)

    fracture at resin composite; (II) fracture at resin composite and coronal tooth structure; (III)

    fracture at resin composite and cervical tooth structure with possible periodontal repair; and

    (IV) root fracture beyond repair.

    2.6 Statistical Analysis

    Statistical analysis was performed by using SPSS 21.0 (SPSS Inc., Chicago, IL,

    USA). The evaluation of optical FBG sensors, cuspal deformation and fracture resistance were

    tested for normal distribution (Shapiro-Wilk, p > 0.05), followed by parametric statistical tests

    and one-way ANOVA with post-hoc Tukey’s test. Data on fracture mode were submitted to

    Chi-square test.

    3 RESULTS

    3.1 Optical FBG Sensors

    Average and standard deviation values of FBG evaluation are presented in Table 2

    and Figure 1. One-way ANOVA revealed statistically significant differences among the

    restorative materials (p < 0.0001). The statistics indicate that the Z100 group presented the

    highest values of stress shrinkage as measured by the optical FBG sensors (p < 0.05). The bulk-

    fill resin groups presented low polymerization stress values, especially the TNC group which

    presented statistically significant difference than FBF group (p < 0.05).

    3.2 Cuspal Deformation

    Average and standard deviation values of cuspal deformation and fracture

    resistance are listed in Table 3. One-way ANOVA analysis demonstrated statistical differences

  • 25

    between the groups (p < 0.03). Z100 group presented higher values of cuspal deformation at

    restoration time and axial compressive compared of others groups (p < 0.01). The cuspal

    deformation until the fracture presented only difference between TNC and FBF groups (p <

    0.05). The fracture resistence of TNC group was higher of all groups, but only statistically

    different of Z100 group (p < 0.01).

    Failure mode distribution is shown in Figure 2 and 3. The Chi-square test (116.000)

    indicates that restorative technique influenced the fracture pattern distribution in the different

    groups (p < 0,0001), with Z100 and FBF groups presenting predominance of type-II fracture

    and TNC and ABF groups presenting type-IV fracture.

    4 DISCUSSION

    The hypothesis predicting that the bulk-fill composite resins generate less

    polymerization contraction stress and reduce cuspal deformation compared to conventional

    composite resins was accepted. The results translate the behavior of the material inside the

    cavity during photopolymerization (Table 2 and 3).

    In dentistry, studies have been introducing optical FBG sensors for assessment of

    dental materials, such as Rajan et al. (2016), who are pioneers in evaluating physical properties

    of different composite resins by using optical fibers, thus providing conclusions on the

    characteristics of these materials during and after photopolymerization. Also, Vinagre et al.

    (2016) evaluated the cuspal deformation by using optical FBG sensors on the tips of cusps

    during photopolymerization of bulk-fill composite resins. This has helped to understand the

    basic characteristics of restorative materials and the influence of polymerization shrinkage

    stress regarding the position of the cusps. However, the present study addressed the

    biomechanical behavior of these materials inside a large MOD cavity, allowing the evaluation

    of the shrinkage stress on the cavity walls and the influenced of the restoration techniques used

    through the optical FBG.

    Restorative materials inside the dental cavity shrink in volume when submitted to

    polymerization, hence producing a tension by contraction, which interacts with their physical

    properties and determine their behavior in the tooth cavity walls [33,34]. The results found

    show that Z100 composite resin has a high shrinkage stress, significantly different of the other

    groups as was reported by Ilie, Kunzelmann and Hickel (2006). The restoration with Z100 was

    performed by using the incremental filling technique in order to minimize the tension created

  • 26

    on the cavity walls. However, this tension was significantly high in the first increment and

    varied strongly in the subsequent ones depending on the position of the increment, also

    demanding a longer chair time for restoration.

    The force distribution found between restorative material and cavity walls

    influenced heavily the prognosis of the restoration procedure, since poorly adapted restorations

    are one of the most frequent causes of failure, thus requiring new restoration. Orlowski,

    Tarczydlo and Chalas (2015) evaluated the marginal adaptation of different bulk-fill resins,

    especially the Filtek and Tetric EvoCream Bulk Fill, and they conclued that the single increment

    technique using this material weakly influence the marginal adjustment and marginal gap.

    The evaluation of cuspal deformation using extensometry, which is a kind of non

    destructive mechanical test, is directly related to the characteristics of the restoration process,

    including restorative and adhesive materials and techniques used, and the size of the cavity

    influences the behavior of the dental structure under forces acting on the adhesive layer [30,34].

    All the specimens were standardized for evaluation of cuspal deformation by using strain

    gauges.

    Bulk-fill resins have physical characteristics such as high translucence to allow light

    passage, different viscosities facilitating their handling, changes in their structure (e.g. volume

    filler) and modifications in the photoinitiator to avoid shrinkage stress [19,23,35]. The resin

    composite TNC have an initiator system fusion, like a camphorquinone conventional and an

    acyl phosphine oxide with Ivocerim® patented system, which it’s more sensitive from

    photopolymerization process requiring lower wavelength comparable of conventional [36]. In

    addition, it’s presented high molecular weight monomers with long chains due to compensate

    the less distance between the formed polymers and particles with low modulus of elasticity [37-

    39].

    In the literature, there are conflicting studies on the behavior of bulk-fill resins

    [12,19]. Our results allow us to state that bulk-fill resins have less polymerization stress than

    the conventional composite ones. The forces generated by cuspal deformation were similar.

    Hence, extensometry and optical FBG sensor tests complement each other, helping confirm the

    behaviour of different materials used for restoring Class II MOD cavity as was observed in the

    study of Lee et al. (2007), which evaluated the cusp deflection of premolars from cavity

    proportion and selected restorative method and, from this, testify a correlation between the cusp

    deflection and stress shrinkage polymerization tests.

  • 27

    In this study was preconized an only kind of adhesive used (Adper Single Bond 2)

    due to standard the adhesive technique, making possible a specific evaluation of material and

    restorative technique to be used, similar of Kim et al. (2015) study.

    The fracture resistance of restorations of Class II MOD cavities was described by

    Eakle (1986), and it can be augmented by composite resin restorations. The results on fracture

    resistance in the present study corroborate the finding reported by Atalay et al. (2016), who

    found similarity between bulk-fill and conventional resins. However, the Z100 resin presented

    low fracture resistance, with predominance of type-II fractures, possibly influenced by its

    behavior during photopolymerization inducing stress accumulation. The higher and significant

    fracture resistance of TNC group is related to its filler content and higher flexural strength

    consequently [38,43].

    5 CONCLUSIONS

    Despite the limitations of this in vitro study, the following conclusions may be

    drawn:

    The bulk fill resin composites inserted in bulk increment of 4 mm depth at short

    photo-activation time generate less shrinkage stress and less cuspal deformation than the

    conventional resin composite.

    The TNC resin composite present low shrinkage stress, cuspal deformation, more

    resistant to fractures however reproduce catastrophic fractures.

    Acknowledgements

    This study was supported by the Gleb Wataghin Institute of Physics (State

    University of Campinas) and Dental Research Center Biomechanics, Biomaterials and Cell

    Biology (Dental School of Federal University of Uberlândia).

    References

    [1] Kwon HJ, Oh YJ, Jang JH, Park JE, Hwang KS, Park YD. The effect of polymerization

    conditions on the amounts of unreacted monomer and bisphenol A in dental composite resins.

    Dent Mater J. 2015;34(3):327-35. doi: 10.4012/dmj.2014-230.

  • 28

    [2] Demarco FF, Collares K, Coelho-de-Souza FH, Correa MB, Cenci MS, Moraes RR,Opdam

    NJ. Anterior composite restorations: A systematic review on long-term survival and reasons for

    failure. Dent Mater. 2015 Oct;31(10):1214-24. doi: 10.1016/j.dental.2015.07.005.

    [3] Demarco FF, Corrêa MB, Cenci MS, Moraes RR, Opdam NJ. Longevity of posterior

    composite restorations: not only a matter of materials. Dent Mater. 2012 Jan;28(1):87-101. doi:

    10.1016/j.dental.2011.09.003.

    [4] Cetin AR, Unlu N, Cobanoglu N. A five-year clinical evaluation of direct nanofilled and

    indirect composite resin restorations in posterior teeth. Oper Dent. 2013 Mar-Apr;38(2):E1-11.

    doi: 10.2341/12-160-C.

    [5] Pashley DH, Tay FR, Breschi L, Tjäderhane L, Carvalho RM, Carrilho M, Tezvergil-

    Mutluay A. State of the art etch-and-rinse adhesives. Dent Mater. 2011 Jan;27(1):1-16. doi:

    10.1016/j.dental.2010.10.016.

    [6] Nahsan FP, Mondelli RF, Franco EB, Naufel FS, Ueda JK, Schmitt VL, Baseggio W.

    Clinical strategies for esthetic excellence in anterior tooth restorations: understanding color and

    composite resin selection. J Appl Oral Sci. 2012 Mar-Apr;20(2):151-6.

    [7] Deliperi S, Bardwell DN. An alternative method to reduce polymerization shrinkage in

    direct posterior composite restorations. J Am Dent Assoc. 2002 Oct;133(10):1387-98. Review.

    Erratum in: J Am Dent Assoc. 2002 Dec;133(12):1614.

    [8] Bönecker M, Tenuta LM, Pucca Junior GA, Costa PB, Pitts N. A social movement to reduce

    caries prevalence in the world. Braz Oral Res. 2013 Jan-Feb;27(1):5-6.

    [9] Da Rosa Rodolpho PA, Donassollo TA, Cenci MS, Loguércio AD, Moraes RR, Bronkhorst

    EM, Opdam NJ, Demarco FF. 22-Year clinical evaluation of the performance of two posterior

    composites with different filler characteristics. Dent Mater. 2011 Oct;27(10):955-63. doi:

    10.1016/j.dental.2011.06.001.

    [10] Ferracane JL. Resin composite--state of the art. Dent Mater. 2011 Jan;27(1):29-38. doi:

    10.1016/j.dental.2010.10.020.

    [11] Casselli DS, Faria-e-Silva AL, Casselli H, Martins LR. Marginal adaptation of class V

    composite restorations submitted to thermal and mechanical cycling. J Appl Oral Sci. 2013 Jan-

    Feb;21(1):68-73.

  • 29

    [12] Benetti AR, Havndrup-Pedersen C, Honoré D, Pedersen MK, Pallesen U. Bulk-fill resin

    composites: polymerization contraction, depth of cure, and gap formation. Oper Dent. 2015

    Mar-Apr;40(2):190-200. doi: 10.2341/13-324-L.

    [13] Fronza BM, Rueggeberg FA, Braga RR, Mogilevych B, Soares LE, Martin AA,

    Ambrosano G, Giannini M. Monomer conversion, microhardness, internal marginal adaptation,

    and shrinkage stress of bulk-fill resin composites. Dent Mater. 2015 Dec;31(12):1542-51. doi:

    10.1016/j.dental.2015.10.001.

    [14] Lee SK, Kim TW, Son SA, Park JK, Kim JH, Kim HI, Kwon YH. Influence of light-curing

    units on the polymerization of low-shrinkage composite resins. Dent Mater J. 2013;32(5):688-

    94.

    [15] Loguercio AD, Reis A, Schroeder M, Balducci I, Versluis A, Ballester RY. Polymerization

    shrinkage: effects of boundary conditions and filling technique of resin composite restorations.

    J Dent. 2004 Aug;32(6):459-70.

    [16] Czasch P, Ilie N. In vitro comparison of mechanical properties and degree of cure of bulk

    fill composites. Clin Oral Investig. 2013 Jan;17(1):227-35. doi: 10.1007/s00784-012-0702-8.

    [17] Roggendorf MJ, Krämer N, Appelt A, Naumann M, Frankenberger R. Marginal quality of

    flowable 4-mm base vs. conventionally layered resin composite. J Dent. 2011 Oct;39(10):643-

    7. doi: 10.1016/j.jdent.2011.07.004.

    [18] Leprince JG, Palin WM, Vanacker J, Sabbagh J, Devaux J, Leloup G. Physico-mechanical

    characteristics of commercially available bulk-fill composites. J Dent. 2014 Aug;42(8):993-

    1000. doi: 10.1016/j.jdent.2014.05.009.

    [19] Orłowski M, Tarczydło B, Chałas R. Evaluation of marginal integrity of four bulk-fill

    dental composite materials: in vitro study. Scientific World Journal. 2015; 2015:701262. doi:

    10.1155/2015/701262.

    [20] Kwon Y, Ferracane J, Lee IB. Effect of layering methods, composite type, and flowable

    liner on the polymerization shrinkage stress of light cured composites. Dent Mater. 2012

    Jul;28(7):801-9. doi: 10.1016/j.dental.2012.04.028.

    [21] Khosravi K, Mousavinasab SM, Samani MS. Comparison of microleakage in Class II

    cavities restored with silorane-based and methacrylate-based composite resins using different

    restorative techniques over time. Dent Res J (Isfahan). 2015 Mar-Apr;12(2):150-6.

    [22] Campos EA, Ardu S, Lefever D, Jassé FF, Bortolotto T, Krejci I. Marginal adaptation of

    class II cavities restored with bulk-fill composites. J Dent. 2014 May;42(5):575-81. doi:

    10.1016/j.jdent.2014.02.007.

  • 30

    [23] El-Damanhoury H, Platt J. Polymerization shrinkage stress kinetics and related properties

    of bulk-fill resin composites. Oper Dent. 2014 Jul-Aug;39(4):374-82. doi: 10.2341/13-017-L.

    [24] Moorthy A, Hogg CH, Dowling AH, Grufferty BF, Benetti AR, Fleming GJ. Cuspal

    deflection and microleakage in premolar teeth restored with bulk-fill flowable resin-based

    composite base materials. J Dent. 2012 Jun;40(6):500-5. doi: 10.1016/j.jdent.2012.02.015.

    [25] Soares CJ, Pizi EC, Fonseca RB, Martins LR. Influence of root embedment material and

    periodontal ligament simulation on fracture resistance tests. Braz Oral Res. 2005 Jan-

    Mar;19(1):11-6.

    [26] Frankenberger R, Krämer N, Appelt A, Lohbauer U, Naumann M, Roggendorf MJ.

    Chairside vs. labside ceramic inlays: effect of temporary restoration and adhesive luting on

    enamel cracks and marginal integrity. Dent Mater. 2011 Sep;27(9):892-8. doi:

    10.1016/j.dental.2011.05.007.

    [27] Anttila EJ, Krintilä OH, Laurila TK, Lassila LV, Vallittu PK, Hernberg RG. Evaluation of

    polymerization shrinkage and hydroscopic expansion of fiber-reinforced biocomposites using

    optical fiber Bragg grating sensors. Dent Mater. 2008 Dec;24(12):1720-7. doi:

    10.1016/j.dental.2008.07.006.

    [28] Urabe I, Nakajima S, Sano H, Tagami J. Physical properties of the dentin-enamel junction

    region. Am J Dent. 2000 Jun;13(3):129-35.

    [29] Bicalho AA, Pereira RD, Zanatta RF, Franco SD, Tantbirojn D, Versluis A, Soares CJ.

    Incremental filling technique and composite material--part I: cuspal deformation, bond

    strength, and physical properties. Oper Dent. 2014 Mar-Apr;39(2):E71-82. doi: 10.2341/12-

    441-L.

    [30] Pereira R, Bicalho AA, Franco SD, Tantbirojn D, Versluis A, Soares CJ. Effect of

    Restorative Protocol on Cuspal Strain and Residual Stress in Endodontically Treated Molars.

    Oper Dent. 2016 Jan-Feb;41(1):23-33. doi: 10.2341/14-178-L.

    [31] Rajan G, Shouha P, Ellakwa A, Bhowmik K, Xi J, Prusty G. Evaluation of the physical

    properties of dental resin composites using optical fiber sensing technology. Dent Mater. 2016

    Sep;32(9):1113-23. doi: 10.1016/j.dental.2016.06.015.

    [32] Vinagre A, Ramos J, Alves S, Messias A, Alberto N, Nogueira R. Cuspal Displacement

    Induced by Bulk Fill Resin Composite Polymerization: Biomechanical Evaluation Using Fiber

    Bragg Grating Sensors. Int J Biomater. 2016;2016:7134283. doi: 10.1155/2016/7134283.

    [33] Ilie N, Kunzelmann KH, Hickel R. Evaluation of micro-tensile bond strengths of composite

    materials in comparison to their polymerization shrinkage. Dent Mater. 2006 Nov;22(7):593-

    601.

  • 31

    [34] Bicalho AA, Valdívia AD, Barreto BC, Tantbirojn D, Versluis A, Soares CJ. Incremental

    filling technique and composite material--part II: shrinkage and shrinkage stresses. Oper Dent.

    2014 Mar-Apr;39(2):E83-92. doi: 10.2341/12-442-L.

    [35] Atalay C, Yazici AR, Horuztepe A, Nagas E, Ertan A, Ozgunaltay G. Fracture Resistance

    of Endodontically Treated Teeth Restored With Bulk Fill, Bulk Fill Flowable, Fiber-reinforced,

    and Conventional Resin Composite. Oper Dent. 2016 Jun 28.

    [36] Zorzin J, Maier E, Harre S, Fey T, Belli R, Lohbauer U, Petschelt A, Taschner M. Bulk-

    fill resin composites: polymerization properties and extended light curing. Dent Mater. 2015

    Mar;31(3):293-301. doi: 10.1016/j.dental.2014.12.010.

    [37] Miletic V, Peric D, Milosevic M, Manojlovic D, Mitrovic N. Local deformation fields and

    marginal integrity of sculptable bulk-fill, low-shrinkage and conventional composites. Dent

    Mater. 2016 Nov;32(11):1441-1451. doi: 10.1016/j.dental.2016.09.011.

    [38] Omran TA, Garoushi S, Abdulmajeed AA, Lassila LV, Vallittu PK. Influence of increment

    thickness on dentin bond strength and light transmission of composite base materials. Clin Oral

    Investig. 2016 Sep 10.

    [39] Rauber GB, Bernardon JK, Vieira LC, Maia HP, Horn F, Roesler CR. In Vitro Fatigue

    Resistance of Teeth Restored With Bulk Fill versus Conventional Composite Resin. Braz Dent

    J. 2016 Jul-Aug;27(4):452-7. doi: 10.1590/0103-6440201600836.

    [40] Lee MR, Cho BH, Son HH, Um CM, Lee IB. Influence of cavity dimension and restoration

    methods on the cusp deflection of premolars in composite restoration. Dent Mater. 2007

    Mar;23(3):288-95.

    [41] Kim RJ, Kim YJ, Choi NS, Lee IB. Polymerization shrinkage, modulus, and shrinkage

    stress related to tooth-restoration interfacial debonding in bulk-fill composites. J Dent. 2015

    Apr;43(4):430-9. doi: 10.1016/j.jdent.2015.02.002.

    [42] Eakle WS. Fracture resistance of teeth restored with class II bonded composite resin. J

    Dent Res. 1986 Feb;65(2):149-53.

    [43] Zhang H, Zhang ML, Qiu LH, Yu JT, Zhan FL. [Comparison of wear resistance and

    flexural strength of three kinds of bulk-fill composite resins]. Shanghai Kou Qiang Yi Xue.

    2016 Jun;25(3):292-5. Chinese.

  • 32

    Tables

  • 33

    Table 2 – Results of FBG evaluation ( µε)

    Z100 TNC FBF ABF

    Average (Standard

    Deviation) 949.1 ± 236.9 A 105.3 ± 205.3 C 525.8 ± 71.2 B 434.8 ± 306.0 BC

    Average followed by the same letter are not statistically different (p > 0.05). n = 5 specimens /

    group.

    Table 3 – Results of cuspal deformation ( µε) and fracture resistance (N)

    Cuspal Deformation

    Restoration Axial Compressive (100N) Fracture Fracture Resistence

    Z100 136.2 ± 41.4 A 89.8 ± 43.1 A 999.1 ± 228.4 AB 1624.2 ± 561.2 B

    TNC 82.6 ± 25.0 B 50.3 ± 31.7 B 1074.5 ± 381.1 A 2925.8 ± 906.6 A

    FBF 87.3 ± 25.9 B 34.1 ± 12.8 B 751.6 ± 243.6 B 1989.9 ± 846.1 AB

    ABF 70.9 ± 11.8 B 34.9 ± 10.4 B 838.1 ± 119.8 AB 2166.1 ± 867.6 AB

    Average followed by the same letter are not statistically different (p > 0.05). n = 10 specimens / group.

  • 34

    Figures

    Figure 1. (a) Typical FBG reflection spectrum. (b) FBG spectra as a function of the applied

    strain. (c) Wavelength shift as a function of the applied strain.

    Figure 2. Percentage of specimens (%) according to fracture pattern classification.

    0%

    10%

    20%

    30%

    40%

    50%

    60%

    70%

    80%

    90%

    100%

    Z100 TNC FBF ABF

    Failu

    re m

    od

    e

    IV (root fracture beyondrepair)

    III (resin composite andcervical tooth structure withpossible periodontal repair)

    II (resin composite andcoronal structure)

    I (resin composite)

  • 35

    Figure 3. SEM representative images of fracture pattern classification. A) Type-1 fracture at

    composite resin (RC); B) Type-2 fracture at composite resin and dentin (D); C) Type-3

    fracture at composite resin and cervical tooth structure, enamel (E) and dentin (D); D) Type-4

    fracture with pulp canal (PC) exposure.

  • 36

    3 CONCLUSÃO

    De acordo com os resultados obtidos e considerando as limitações das metodologias

    utilizadas, pôde-se concluir que:

    1. Resinas compostas bulk fill apresentam menor tensão de contração e deformação

    de cúspide durante a fotopolimerização de um único incremento com 4 mm de espessura em

    relação à resina composta convencional.

    2. A resina composta TNC apresenta menor tensão de contração de polimerização,

    menor deformação de cúspide, além de ser mais resistente à fraturas, contudo, apresenta maior

    frequência de fraturas catastróficas.

    3. O padrão de fratura das resinas compostas quando induzidas por forças

    compressivas é diretamente dependente do material resinoso utilizado.

  • 37

    REFERÊNCIAS

    Al-Fakih E, Abu Osman NA, Mahamd Adikan FR. The use of fiber Bragg grating sensors in

    biomechanics and rehabilitation applications: the state-of-the-art and ongoing research topics.

    Sensors (Basel). 2012 Sep 25;12(10):12890-926. doi: 10.3390/s121012890.

    Anatavara S, Sitthiseripratip K, Senawongse P. Stress relieving behaviour of flowable

    composite liners: A finite element analysis. Dent Mater J. 2016;35(3):369-78. doi:

    10.4012/dmj.2015-204

    Atalay C, Yazici AR, Horuztepe A, Nagas E, Ertan A, Ozgunaltay G. Fracture Resistance of

    Endodontically Treated Teeth Restored With Bulk Fill, Bulk Fill Flowable, Fiber-reinforced,

    and Conventional Resin Composite. Oper Dent. 2016 Jun 28.

    Benetti AR, Havndrup-Pedersen C, Honoré D, Pedersen MK, Pallesen U. Bulk-fill resin

    composites: polymerization contraction, depth of cure, and gap formation. Oper Dent. 2015

    Mar-Apr;40(2):190-200. doi: 10.2341/13-324-L.

    Bicalho AA, Pereira RD, Zanatta RF, Franco SD, Tantbirojn D, Versluis A, Soares CJ.

    Incremental filling technique and composite material--part I: cuspal deformation, bond

    strength, and physical properties. Oper Dent. 2014 Mar-Apr;39(2):E71-82. doi: 10.2341/12-

    441-L.

    Bönecker M, Tenuta LM, Pucca Junior GA, Costa PB, Pitts N. A social movement to reduce

    caries prevalence in the world. Braz Oral Res. 2013 Jan-Feb;27(1):5-6.

    Campos EA, Ardu S, Lefever D, Jassé FF, Bortolotto T, Krejci I. Marginal adaptation of class

    II cavities restored with bulk-fill composites. J Dent. 2014 May;42(5):575-81. doi:

    10.1016/j.jdent.2014.02.007.

    De acordo com as normas da UNICAMP/FOP, baseadas na padronização do International Committee of Medical

    Journal Editors – Vancouver Group. Abreviatura dos periódicos em conformidade com o PubMed.

  • 38

    Casselli DS, Faria-e-Silva AL, Casselli H, Martins LR. Marginal adaptation of class V

    composite restorations submitted to thermal and mechanical cycling. J Appl Oral Sci. 2013

    Jan-Feb;21(1):68-73.

    Cetin AR, Unlu N, Cobanoglu N. A five-year clinical evaluation of direct nanofilled and

    indirect composite resin restorations in posterior teeth. Oper Dent. 2013 Mar-Apr;38(2):E1-

    11. doi: 10.2341/12-160-C.

    Czasch P, Ilie N. In vitro comparison of mechanical properties and degree of cure of bulk fill

    composites. Clin Oral Investig. 2013 Jan;17(1):227-35. doi: 10.1007/s00784-012-0702-8.

    Da Rosa Rodolpho PA, Donassollo TA, Cenci MS, Loguércio AD, Moraes RR, Bronkhorst

    EM, et al. 22-Year clinical evaluation of the performance of two posterior composites with

    different filler characteristics. Dent Mater. 2011 Oct;27(10):955-63. doi:

    10.1016/j.dental.2011.06.001.

    Deliperi S, Bardwell DN. An alternative method to reduce polymerization shrinkage in direct

    posterior composite restorations. The Journal of the American Dental Association. 2002;

    133(10): 1387-1398.

    Demarco FF, Collares K, Coelho-de-Souza FH, Correa MB, Cenci MS, Moraes RR, et al.

    Anterior composite restorations: A systematic review on long-term survival and reasons for

    failure. Dent Mater. 2015 Oct;31(10):1214-24. doi: 10.1016/j.dental.2015.07.005.

    Demarco FF, Corrêa MB, Cenci MS, Moraes RR, Opdam NJ. Longevity of posterior

    composite restorations: not only a matter of materials. Dent Mater. 2012 Jan;28(1):87-101.

    doi: 10.1016/j.dental.2011.09.003.

    Dominguez JA, Bittencourt BF, Farago PV, Pinheiro LA, Gomes OMM. Shrinkage stress of

    resin composites: effect of material composition - sistematic review. Braz Dent Sci 2014

    Jul/Sep;17(3).

    El-Damanhoury H, Platt J. Polymerization shrinkage stress kinetics and related properties of

    bulk-fill resin composites. Oper Dent. 2014 Jul-Aug;39(4):374-82. doi: 10.2341/13-017-L.

  • 39

    Ferracane JL. Resin composite--state of the art. Dent Mater. 2011 Jan;27(1):29-38. doi:

    10.1016/j.dental.2010.10.020.

    Fronza BM, Rueggeberg FA, Braga RR, Mogilevych B, Soares LE, Martin AA, et al.

    Monomer conversion, microhardness, internal marginal adaptation, and shrinkage stress of

    bulk-fill resin composites. Dent Mater. 2015 Dec;31(12):1542-51. doi:

    10.1016/j.dental.2015.10.001.

    Kalliecharan D, Germscheid W, Price RB, Stansbury J, Labrie D. Shrinkage stress kinetics of

    Bulk Fill resin-based composites at tooth temperature and long time. Dent Mater. 2016

    Nov;32(11):1322-1331. doi: 10.1016/j.dental.2016.07.015.

    Kwon HJ, Oh YJ, Jang JH, Park JE, Hwang KS, Park YD. The effect of polymerization

    conditions on the amounts of unreacted monomer and bisphenol A in dental composite resins.

    Dent Mater J. 2015;34(3):327-35. doi: 10.4012/dmj.2014-230.

    Lee SK, Kim TW, Son SA, Park JK, Kim JH, Kim HI, et al. Influence of light-curing units on

    the polymerization of low-shrinkage composite resins. Dent Mater J. 2013;32(5):688-94.

    Leprince JG, Palin WM, Vanacker J, Sabbagh J, Devaux J, Leloup G. Physico-mechanical

    characteristics of commercially available bulk-fill composites. J Dent. 2014 Aug;42(8):993-

    1000. doi: 10.1016/j.jdent.2014.05.009.

    Loguercio AD, Reis A, Schroeder M, Balducci I, Versluis A, Ballester RY. Polymerization

    shrinkage: effects of boundary conditions and filling technique of resin composite

    restorations. J Dent. 2004 Aug;32(6):459-70.

    Moorthy A, Hogg CH, Dowling AH, Grufferty BF, Benetti AR, Fleming GJ. Cuspal

    deflection and microleakage in premolar teeth restored with bulk-fill flowable resin-based

    composite base materials. J Dent. 2012 Jun;40(6):500-5. doi: 10.1016/j.jdent.2012.02.015.

  • 40

    Nahsan FP, Mondelli RF, Franco EB, Naufel FS, Ueda JK, Schmitt VL, et al. Clinical

    strategies for esthetic excellence in anterior tooth restorations: understanding color and

    composite resin selection. J Appl Oral Sci. 2012 Mar-Apr;20(2):151-6.

    Orłowski M, Tarczydło B, Chałas R. Evaluation of marginal integrity of four bulk-fill dental

    composite materials: in vitro study. Scientific World Journal. 2015; 2015:701262. doi:

    10.1155/2015/701262.

    Ozakar-Ilday N, Zorba YO, Yildiz M, Erdem V, Seven N, Demirbuga S. Three-year clinical

    performance of two indirect composite inlays compared to direct composite restorations. Med

    Oral Patol Oral Cir Bucal. 2013 May 1;18(3):e521-8.

    Park JW, Ferracane JL. Measuring the residual stress in dental composites using a ring slitting

    method. Dent Mater. 2005 Sep;21(9):882-9.

    Park JW, Ferracane JL. Residual stress in composites with the thin-ring-slitting approach. J

    Dent Res. 2006 Oct;85(10):945-9.

    Pashley DH, Tay FR, Breschi L, Tjäderhane L, Carvalho RM, Carrilho M, et al. State of the

    art etch-and-rinse adhesives. Dent Mater. 2011 Jan;27(1):1-16. doi:

    10.1016/j.dental.2010.10.016.

    Pereira R, Bicalho AA, Franco SD, Tantbirojn D, Versluis A, Soares CJ. Effect of Restorative

    Protocol on Cuspal Strain and Residual Stress in Endodontically Treated Molars. Oper Dent.

    2016 Jan-Feb;41(1):23-33. doi: 10.2341/14-178-L.

    Rajan G, Shouha P, Ellakwa A, Bhowmik K, Xi J, Prusty G. Evaluation of the physical

    properties of dental resin composites using optical fiber sensing technology. Dent Mater. 2016

    Sep;32(9):1113-23. doi: 10.1016/j.dental.2016.06.015.

    Roggendorf MJ, Krämer N, Appelt A, Naumann M, Frankenberger R. Marginal quality of

    flowable 4-mm base vs. conventionally layered resin composite. J Dent. 2011

    Oct;39(10):643-7. doi: 10.1016/j.jdent.2011.07.004.

  • 41

    Umesh S, Padma S, Asokan S, Srinivas T. Fiber Bragg Grating based bite force measurement.

    J Biomech. 2016 Jul 1. pii: S0021-9290(16)30732-1. doi:10.1016/j.jbiomech.2016.06.036.

    Van Dijken JW, Pallesen U. Posterior bulk-filled resin composite restorations: A 5-year

    randomized controlled clinical study. J Dent. 2016 Aug;51:29-35.

    doi:10.1016/j.jdent.2016.05.008.

    Vinagre A, Ramos J, Alves S, Messias A, Alberto N, Nogueira R. Cuspal Displacement

    Induced by Bulk Fill Resin Composite Polymerization: Biomechanical Evaluation Using

    Fiber Bragg Grating Sensors. Int J Biomater. 2016;2016:7134283. doi:

    10.1155/2016/7134283.

    Witzel MF, Braga RR, Ballester RY, Lima RG. Influence of specimen dimensions on nominal

    polymerization contraction stress of a dental composite. J. of the Braz. Soc. of Mech. Sci. &

    Eng. 2005 Jul-Sept 27(3):283-287.

    Yeh CH, Chow CW, Wu PC, Tseng FG. A simple fiber Bragg grating-based sensor network

    architecture with self-protecting and monitoring functions. Sensors (Basel). 2011;11(2):1375-

    82. doi: 10.3390/s110201375.

  • 42

    APÊNDICE 1

    Metodologia Ilustrada

    Figura 1. Preparo das amostras. A) Seleção de molares hígidos; B) Simulação do

    ligamento periodontal 2 mm abaixo da junção amelocementária com cera 7; C) Posicionamento

    do dente à haste de um delineador protético; D) Centralização de uma película radiográfica com

    perfuração central ao dente; E) Posicionamento de um tubo de PVC para inclusão da porção

    radicular em resina de poliestireno; F) Inclusão do dente na resina de poliestireno; G) Remoção

    dos dentes nos alvéolos artificiais e eliminação da cera 7 na raiz do dente; H) Realização da

    cavidade MOD em máquina de preparo; I) Cavidade MOD confeccionada; J) Inclusão do

    poliéter Impregum.

  • 43

    Figura 2. Avaliação da tensão de contração de polimerização dos materiais

    restauradores pelos sensores de Bragg. A) Condicionamento com ácido fosfórico a 37% no

    esmalte por 15 segundos; B) Condicionamento com ácido fosfórico sobre a dentina por 15

    segundos e no esmalte por mais 15 segundos; C) Lavagem abundante com água por 30

    segundos; D) Controle da umidade mantendo a dentina levemente úmida; E) Fibra óptica

    esticada com a presença dos sensores de Bragg na região azul; F) Posicionamento dos sensores

    de Bragg na região amelodentinária justaposto à parede vestibular da cavidade MOD; G)

    Aplicação do adesivo Adper Single Bond 2 seguido da volatilização do solvente por 5 segundos;

    H) Fotopolimerização do adesivo por 40 segundos; I) Fibra óptica posicionada na parede

    vestibular da cavidade MOD com o sensor de Bragg aderido à parede com adesivo fotoativado.

  • 44

    Figura 3. Teste de extensometria para avaliação da deformação de cúspide. A)

    Aplicação de ácido fosfórico por 30 segundos na região determinada para instalação do

    extensômetro; B) Posicionamento do extensômetro na base da cúspide vestibular e

    lingual/palatina ao nível da parede gengival das caixas proximais da cavidade MOD com cola

    a base de cianoacrilato; C) Fotopolimerização das restaurações conforme o protocolo dos

    grupos de tratamento e avaliação da deformação de cúspide pelos extensômetros; D) Ligação

    dos extensômetros à máquina analisadora por meio de pontos de solda; E) Avaliação da

  • 45

    deformação de cúspide durante carregamento de 100N e até a fratura em máquina de ensaio

    universal.

    Figura 4. Protocolo restaurador dos grupos com resina convencional e resina bulk

    fill. A) Inserção de forma incremental do G Z100; B C e D) Inserção de um único incremento

    de 4mm dos grupos TNC, FBF e ABF, respectivamente.

  • 46

    ANEXO 1

  • 47

    ANEXO 2: Dental Materials