clinical and scientific aspect of inlay fixed partial dentures

129
1 UNIVERSITY OF SIENA SCHOOL OF DENTAL MEDICINE PHD PROGRAM: “DENTAL MATERIALS AND CLINICAL APPLICATIONS” Ph D THESIS OF: Carlo Monaco TITLE “Clinical and scientific aspect of Inlay Fixed Partial Dentures”

Upload: others

Post on 12-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

1

UNIVERSITY OF SIENA

SCHOOL OF DENTAL MEDICINE

PHD PROGRAM:

“DENTAL MATERIALS AND CLINICAL APPLICATIONS”

Ph D THESIS OF:

Carlo Monaco

TITLE

“Clinical and scientific aspect of Inlay Fixed Partial Dentures”

2

ACCADEMIC YEAR 2004/2005 December 2005

Siena Italy

Committee:

Promoter Prof. Marco Ferrari

Co-Promoter Prof. R. Scotti

Prof. xxxxxxxxxxxx

Prof. xxxxxxxxxxxx

Prof. xxxxxxxxxxxx

Prof. xxxxxxxxxxxx

TITLE

“Clinical and scientific aspect of Inlay Fixed Partial Dentures” _____________________________________________________________

CANDIDATE

Carlo Monaco

December 2005

3

CONTENTS

Chapter 1: General introduction

1.1 Tooth structure removal associated with various preparation designs

1.2 Metal-free inlay retainer restorations

1.3 Indication and contraindications of inlay-fixed partial denture

Chapter 2: The use of fiber reinforced composites in dentistry

2.1 Fiber-reinforced composite systems

2.2 Properties of the fibers and polymer matrices

2.3 Impregnation of the fibers

2.4 Quantity of fibers

2.5 Direction of the fibers

2.6 Position of fibers

2.7 Water sorption of FRC matrix

Chapter 3: Marginal adaptation of IFPDs

3.1 Marginal adaptation of three partial bridges made with different structure

material.

Chapter 4: Criteria for selecting the materials for IFPDs

4.1 Fracture strength of three partial bridges made with different structure material.

Chapter 5: laboratory process for high volume fiber framework

5.1 Fiber reinforced composite with a high volume framework: a technical

procedure.

Chapter 6: Different structure of the framework

6.1 Clinical Evaluation of Fiber-Reinforced Composite IFPDs.

Chapter 7: adhesive procedures

7.1 Inlay Bridge With a New Microfilled Composite: A Clinical Report

4

Chapter 8: Clinical trial

8.1 Randomized controlled trial of Fiber-Reinforced Composite Inlay Fixed Partial

Dentures: two-year results.

Chapter 9: Alternative materials as regards FRC

9.1 Fatigue test in shear: its effect on bond of a glass-infiltrated alumina ceramic to

human dentin, using different luting procedures.

Chapter 10 Other clinical application of FRC

11.1 Clinical evaluation of teeth restored with quartz fiber-reinforced epoxy resin

posts.

Chapter 11 Conclusions

Summary

5

Chapter 1 General Introduction

When missing tooth structure or teeth are replaced, minimal biologic risk should be

involved to reestablish function and esthetics. To proven reliability and durability of

complete-crown metal ceramics made them the method of choice for posterior single-

tooth restorations and fixed partial denture (FPD). However, this restoration required

considerable reduction of tooth structure. The increased use of the adhesive

technique and preservation of dental tissues have greatly impacted conservative tooth

preparation design. The development of fibre-reinforced composite (FRC)

technology and all-ceramic systems has opened the potential for fabrication of metal-

free restorations with durability and good aesthetics.

This thesis contains a study on several different basic and clinical aspects related to

the use of inlay-fixed partial dentures made with fiber-reinforced composites and all

ceramic systems.

Starting from the assessment of the differences between the amount of tooth structure

removed for conventional preparation and various innovative designs for fixed

prosthodontics, the next step was to analyse the different materials that can be used

when missing tooth must be replaced. Inlay-fixed partial dentures and dental

implants are the true alternatives to the conventional metal-ceramic three-unit

bridges; for these reason an overview regarding the properties advantages and

disadvantages of fiber-reinforced materials is presented. As actually different fiber-

reinforced composites are available on the market, it is important for the clinician to

know the properties of each system to select the more appropriate for the specific

clinical application.

The first objective of this thesis was to evaluate before and after fatigue the marginal

adaptation of inlay fixed partial dentures made with different materials and establish

a connection between the quality of continuous/non-continuous margins and the

mechanical properties of different materials. The second was indeed to conduct a

study to assess the fracture strength and the dye penetration after fatigue of one fiber-

reinforced composite and two all-ceramic systems, and to verify the existence of a

correlation between the mechanical resistance and the microleakage. Another goal of

this thesis was in fact to evaluate if and how different methods of positioning of the

6

fibers for the framework can increase the fracture strength and reduced the flexibility

of the bridges.

The design of the fiber framework is an important prerequisite to obtain a durable

clinical success when using fiber-reinforced composite. Another step of this thesis

was to describe the technical procedure to obtain a framework with a high volume of

fiber; for these reason a clinical study comparing the survival rate of inlay fixed

partial dentures made with different framework design is presented.

Bonding procedures represent the goal for the term of partial restorations. The next

step of this thesis was to describe the luting procedures and the surface treatment for

the inlay bridges and to compare the clinical performances and the post-operative

sensibility of three- and two-step adhesive systems after two-year observation period.

Alternative materials to the metal-ceramic restorations went through rapid

developments in the last few years, in particular alumina- and zirconia based

ceramics. Both these materials represent the future alternatives of the fiber-reinforced

composite but their clinical applications in partial restorations are still limited. The

next steps of this thesis were the analysis of the bond of a glass-infiltrated alumina

ceramic to human dentin, using different luting procedures.

Finally the use of fiber-reinforced composite in the reconstruction of the

endodontically treated teeth is examined with according to the aim of the minimal

intervention philosophy.

1.1 Tooth structure removal associated with various preparation designs

The introduction of more invasive complete crown preparation for metal- and all-

ceramic crowns has been correlated with an increase in pulpal complications since

these restorations require considerable reduction of tooth structure (Creugers et al

1994). For a metal-ceramic shoulder preparation, a facial tooth reduction of about 1.3

to 1.5 mm and an occlusal reduction of 2.0 mm are recommended (McLean JW 1980,

Rosenstiel et al 1995). In 1966, only 0.4% to 2% radiographic periapical pathologies

were found (Ericsson et al 1966), whereas in 1970, 2.9% was reported (Schwartz et

al 1970), and about 10 years later up to 4.0% periapical pathologies were detected

(Kerschbaum et at 1981). These results are explained by the use of air turbines () and

more invasive shoulder or chamfer preparations compared to the feather-edge design

7

used in the 1960s and 1970s (Klötzer 1984). A lower number of endodontic

complications are associated with less invasive preparations. In a literature review,

inlay restorations at 10 years showed a lower rate of loss of pulpal vitality (5.5%)

compared to complete crowns (14.5%) (Kerschbaum et al 1981). The mechanical

reliability and broad range of indications have made complete crowns the preferred

denture retainer. However, wing—shaped retainers with retentive elements such as

grooves made of metal have demonstrated a remarkable long-term success rate if the

clinical protocol is followed carefully (Creugers et al 1992). The gravimetric analysis

(Edelhoff et al 2002) showed that for a metal-ceramic crown retainer preparation,

almost eight times more tooth structure must be removed compared to an adhesive

wing-and groove attachment for a resin-bonded cast-metal fixed partial denture. The

“new” half-crown preparation assigned for all ceramic fixed partial dentures (FPDs)

required a similar amount of tooth structure removal as the onlay and cost

approximately half of the tooth structure of a complete crown design. The percentage

of tooth structure removal associated with the different preparation designs for a

mandibular premolar was 19.3% for mesial/distal occlusal inlay without transverse

ridge or central groove, 30.4% for mesial/distal occlusal inlay with transverse ridge

or central groove and 75.9% for mete-ceramic complete crown. Similar percentages

of tooth structure removal were found for the same kind of preparation in mandibular

molar (19.3%, 25.5%, 73.1%).

The inclusion of enamel promotes a superior bond over dentin, lower post-

cementation sensitivity, improved support of the materials used for the restorations,

and reduced endodontic intervention.

The positive influence of tooth structure preservation on the life expectancy of the

pulp was reported in the literature. For cast-metal resin- bonded FPDs, a 0.13% rate

of loss of pulpal vitality up to 5 years was reported, compared to 9.1% for complete

crown abutments in the same period (Paszyna et al 1990).

1.2 Metal-free inlay retainer restorations

For the past 30 yr, some dentists have avoided the use of full coverage retainers for

fixed partial dentures in order to conserve sound tooth substance. Generally, metal-

reinforced systems are the materials of choice for fabricating posterior fixed partial

dentures because of their reliability and durability. Inlay-retained FPDs made of

8

metal alloys are been usually seated using the conventional cementation technique

and cements (Kopp 1970). Before adhesive techniques were introduced to restorative

dentistry, conventionally cemented partial crowns or inlays, made of cast gold, were

used instead of full coverage crowns to retain a pontic (Boitel 1969). A common

problem was the loss of retention of a retainer, with subsequent secondary caries

development (Roberts 1970). As a result, more effective intracoronal retention with

the help of boxes, grooves, and pins was demanded (Weinberg 1970). These

solutions, however, mitigated the advantage of minimal invasiveness compared with

complete-crown retainers. In the 1980s, adhesive techniques allowed the luting of

metallic frameworks to dental enamel by using metal retainer wings made of cast

gold or non-precious metal (Rochette 1973, Livaditis 1983). Inadequately retentive

preparation shapes and insufficient stability of the metal framework were perceived

to have been contributing factors. After initially frequent losses of retention, more

defined and retentive preparations, along with improved adhesives, led to acceptable

retention rates, especially in anterior teeth (Rammelsberg et al 1993). The aesthetic

limitations caused by the metallic framework remained a problematic issue. The dark

framework on the oral surfaces of abutment teeth eliminated translucency and gave

the teeth a greyish appearance (Livaditis 1983) Restorations made of metal alloys are

characterized by certain basic disadvantages. These base metal components that form

on the surface of the alloy during the metal-ceramic fusing process may have a

negative effect on the adjacent soft tissue. In addition, the opaque, darkish

appearance created by certain metal denture retainers in the abutment teeth is

considered to be unattractive. Partial preparations like inlays, onlays or partial

crowns are recommended as retainers for short-span FPDs in caries-resistant

dentitions. In addition to facilitating superior periodontal health, partial retainers

enable preservation of healthy tooth structure. The combination of highly translucent

prosthodontic materials and resin composite cements has enhanced the use of the

adhesive technique and launched a new era of restorative treatment options with

promising initial clinical results (Sorensen et al 1999). New in vitro findings and a

better understanding of stress formation in fiber-reinforced composite (FRC)

(Vallittu 1996, Freilich et al 2004) and in all-ceramic restorations led to less invasive

preparations extended to existing systems. There has been limited use and no

published clinical data of all-ceramic posterior FPDs retained either by wings or

9

inlays, mostly because of the low strength, the strength scatter, and the time-

dependent strength decrease of ceramics owing to slow crack growth (Fischer et al

2003). The reduced invasiveness of these resin-bonded inlay-retained FPDs makes

them an appealing alternative to conventional preparations in cases where the

residual dentition exhibits low caries activity. Metal-free materials such as fiber-

reinforced composites or high strength pressed ceramics exhibit outstanding

corrosion resistance. The esthetics properties of these systems must be attributed to

the high translucency of the materials and the fact than the restorations are entirely

fabricated of tooth-coloured materials thereby achieving a high degree of light

transmission. However, restorations made of these materials are not as strong as

those that are metal-supported because of their particular mechanical properties. To

achieve adequately strong dental restorations, therefore, certain modifications are

necessary in the preparations fabrications, and cementation methods. The preparation

geometry on an inlay retainer offers favourable prerequisites for the adhesive

cementations technique. The preparation is usually surrounded by dental enamel, and

the location of the preparation margin allows a rubber dam to be placed to ensure

complete isolation. Adhesive cementation could offer one of the most effective ways

of countering the loss of retention, which is one of the most frequent causes of failure

of conventional inlay-retained fixed partial dentures.

1.3 Indication and contraindications of inlay-fixed partial denture

The indications have to be strictly observed because of the special properties of the

metal-free materials. As a result, careful assessment and planning prior to beginning

the prosthodontic treatment measures are requisite. Furthermore, the following

prerequisites must be met if the successful results are to be achieved with metal free-

inlay-retained FPDs.

1. Good oral hygiene

2. Low susceptibility to caries

3. Parallel alignment of abutment teeth

4. Immobility of the abutment teeth

5. Minimum height of abutment teeth ≥ 5mm (connector thickness)

10

6. Maximum mesiodistal extension of the interdental gap of 9 mm (width of

premolar) if pressed ceramic is used and 12 mm (width of molar) if fiber-

reinforced composite materials are used.

Severe parafunctions, short clinical crowns (<5mm) and extensive defects of the

clinical crown, as well as the loosening of teeth because of factors related to the

periodontium, have been established as contraindications. The cusps of

endodontically pre-treated teeth are included in the preparation to protect them.

Metal-free inlay-retained fixed partial dentures must be adhesively cemented because

of the primary friction compared with metal-supported systems (Edelhoff et al 2001).

Existing therapy-resistant periodontopathologic complains as well as allergies to the

components of dentin adhesives or luting composites, therefore, must be classified as

absolute contraindications. Gingival bleeding could compromise the adhesive bond

between the resin and the prepared tooth. In preparation for adhesive cementation,

therefore, all signs of periodontal inflammation should be eliminated.

In addition to a thorough intraoral examination, radiographs (proximal caries,

periodontum) of the designed abutment teeth and irreversible hydrocolloid

impressions recommended for evaluating these factors. Special attention must be

paid to generalized wear facets, the position of the antagonist contacts, existing

hyperbalances, the length of the clinical crown, the pontic span, and the alignment of

the abutment teeth. In addition, canine guidance must be ensured to protect the inlay-

retained fixed partial denture from torsional stress. If it is not established, its

reconstruction during the restorative procedures should be considered.

11

References

Boitel RH. Pin abutment for crown and bridge work. Dtsch Zahnarztl Z 1969;

24:705–707.

Creugers NH, Kayser AF, van 't Hof MA. A meta-analysis of durability data

on conventional fixed bridges. Community Dent Oral Epidemiol. 1994 Dec;22:448-

52.

Creugers NH, Kayser AF, Van't Hof MA. A seven-and-a-half-year survival

study of resin-bonded bridges. J Dent Res. 1992 Nov;71:1822-5.

Edelhoff D, Sorensen JA. Tooth structure removal associated with various

preparation designs for posterior teeth Int J Periodontics Restorative Dent.

2002;22:241-9.

Edelhoff D, Spiekermann H, Yildirim M. Metal-free inlay-retained fixed

partial dentures. Quintessence Int. 2001 Apr;32(4):269-81.

Ericsson S, Hedegard B, Wennstrom A. Roentgenographic study of vital

abutment teeth. J Prosthet Dent 1966;16:981-987.

Fischer H, Weber M, Marx R. Lifetime prediction of allceramic bridges by

computational methods. J Dent Res 2003;82: 238–242.

Freilich MA, Meiers JC. Fiber-reinforced composite prostheses. Dent Clin

North Am. 2004, 448:545-62.

Kerschbaum T, Voss R. Practical efficacy of crowns and inlays Dtsch

Zahnarztl Z. 1981 Apr;36:243-9.

Klötzer WT. Die traumatische Schadigung der pulpa bei der Uberkronung.

Dtsch Zahnarztl 1984;39:791-794

Kopp EN. Partial veneer retainers. J Prosthet Dent. 1970;23:412-9.

Livaditis GJ. Etched metal resin-bonded restorations: principles in retainer

design. Int J Periodontics Restorative Dent 1983; 3: 34–47.

McLean JW. The cast metal-ceramic crown. In: The science and art of dental

ceramics. Chicago: Quintessence, 1980:202.

Paszyna C, Kerschbaum T, Marinello CP, Pfeiffer P. Clinical long-term

results with bonded bridges Dtsch Zahnarztl Z. 1990 Jul;45:406-9.

Rammelsberg P, Pospiech P, Gernet W. Clinical factors affecting adhesive

12

fixed partial dentures: a 6-year study. J Prosthet Dent 1993; 70: 300–307.

Roberts DH. The failure of retainers in bridge prostheses. An analysis of

2,000 retainers. Br Dent J 1970; 128: 117–124

Rochette AL. Attachment of a splint to enamel of lower anterior teeth. J

Prosthet Dent 1973; 30: 418–423.

Rosenstiel SF, Land MF, Fujimoto J. The contemporary Fixed

Prosthodontics. The metal-ceramic crown preparation, ed 2. St.Louis :Mosby Year

Book, 1995:180-193.

Schwartz NL, Whitsett LD, Berry TG, Stewart JL. Unserviceable crowns and

fixed partial dentures: lifespan and causes for loss of serviceability. Am J Dent

1970;81:1395-1401.

Sorensen JA, Cruz M, Mito WT, Raffeiner O, Meredith HR, Foser HP. A

clinical investigation on three-unit fixed partial dentures fabricated with a lithium

disilicate glass-ceramic. Pract Periodontics Aesthet Dent. 1999 Jan-Feb;11:95-106.

Vallittu PK. A review of fiber-reinforced denture base resins. J Prosthodont

1996;5(4):270—6.

Weinberg LA. Vertical nonparallel pin-inlay fixed partial prosthesis.J

Prosthet Dent. 1970;23:420-33.

13

Chapter 2: The use of fiber reinforced composites in dentistry

Fibre-reinforced materials combine the basically different mechanical properties of

fibres and a matrix, in which the fibres are embedded. The fibres demonstrate high

tensile strength, a high tensile modulus, and low shear strength, while the matrix is

characterized by high toughness. In an optimum fibre-reinforced material, the tensile

strength of the fibre is combined with the high toughness of the matrix. Fibre-

reinforced technology is used wherever high stress occurs and low weight is required,

such as in the aeronautical and shipbuilding industries.

Low weight combined with high strength is also required in removable denture

prosthetics, for which PMMA resins have proved to be particularly suitable due to

their resistance to the oral environment. Since complete dentures may fracture, glass-

fibre reinforcements had been discussed by dental interest groups for decades.

(Grotsch, 1965a; Grotsch, 1965b; Mc Creight, 1967). Research mainly focused on

the reinforcement of PMMA denture base materials by means of fibres (Vallittu,

1996). Most scientists found that increased mechanical strength values can be

achieved by means of fibre reinforcement (Vallittu, 1996), with the fracture

resistance enhancing with an increasing fibre content (Vallittu et al. 1994).

2.1 Fiber-reinforced composite systems

Glass fibre-reinforced composites (FRC) were introduced to dentistry in the late

1990s and were advertised as a universal aesthetic material for nearly every dental

indication. Several in vitro studies confirmed good material properties and good

marginal adaptation (Behret et al., 1999; Körber et al., 1996; Göhring et al., 2001,

Goldberg et al., 1992; Behr et al., 1999; Karmaker et al., 1997).

Composite materials are a combination of two or more distinct components forming

a new material with enhanced properties. While many combinations exist, the most

common composites in engineering are composed of strong fibers held by a binder or

matrix. Unlike traditional materials, the properties of composites can be designed

simultaneously with structural aspects. This allows composite designers to

manipulate material properties by changing fiber orientation, fiber content, and

geometry. Additionally, the most common types of matrix materials are polymers

14

(Barbero, 1998). Attempts have been made to reinforce dental polymers with several

types of fibers for various treatment modalities during the past 30 years. Studies have

tested polyethylene fibers (Ladizesky et al., 1992), carbon/ graphite fibers (Kilfoil et

al., 1983; Malquarti et al., 1990; Ruyter et al., 1986), or glass fibers (Goldberg et al.,

1992, Imai et al., 1999; Meiers et al., 2000; Vallittu et al., 1996). There exist

potential applications for fiber-reinforced composites (FRC) in prosthodontics,

periodontics, and orthodontics. Several in vitro studies have been conducted to find

out and understand the factors influencing dental FRC properties (Vallittu et al.,

1994; Viguie et al., 1994; Behr et al., 2000; Vallittu et al., 1998; Vallittu et al., 2000;

Nohrstrom et al., 2000). Important factors influencing the mechanical properties of

FRCs include:

(1) inherent material properties of fibers and polymer matrices,

(2) fiber surface treatment (sizing) and impregnation of fibers with resin,

(3) quantity of fibers (Lassila et al., 2002),

(4) direction of fibers

(5) position of fibers (Vallittu, 2002; Chung et al 1998; Ellakwa et al., 2001)

(6) water sorption of FRC matrix (Lassila et al., 2002).

The main FRCs are represented by systems with different characteristics. Glass

fiber–reinforced systems (Vectris; Ivoclar-Vivadent, Schaan, FL and FibreKor;

Jeneric/Pentron, Wallingford, CT) use continuously oriented fibers preimpregnated

with monomers ready for heat or light curing (“prepregs”). FibreKor prepregs are

unidirectional and adapted manually. Vectris uses a vacuum/pressure device to shape

the framework. The Vectris framework consists of 3 different prepregs that can be

distinguished by the fiber orientation; prepregs with parallel fibers are called

“pontic,” those with a 45-degree alignment are called “single,” and those with a 90-

degree alignment are the “frame” prepregs. A third prepreg system (EverStick,

Turku, Finland) comprises glass fibers preimpregnated with thermoplastic polymers.

Another system (Connect-Band; SDS Belle, Orange, CA) consists of woven

“plasma-etched” polyethylene fibers that must be impregnated by the user before

manual adaptation. Fixation of the fibers in the matrix occurs only by mechanical

means, and perfect impregnation depends on the skill of the user.

15

2.2 Properties of the fibers and polymer matrices

The composition of glass fibers used for dental applications varies. Continuous fibers

for polymer-glass fiber composite are usually made of alkali-free glass (up to 1%

Na2O + K2O) know as E-glass (electrical glass). E-glass is based on the SiO2-Al2-O3-

CaO-MgO system, which has good glass-forming ability. Because of the high

calcium oxide (CaO) content, glass similar to this composition shows poor resistance

to acidic solutions. For this reason, the composition of E-glass is modified by

introducing boron oxide (B2O3) and by decreasing the CaO content (Hlavác, 1983).

The composition for E-glass is presented in table 1. Four other types of glass used in

polymer-fiber composites, S-glass and R-glass (both high-tensile-strength glass),

acid-resistant, and alkaline-resistant glass, are also included. Glass composition

influences the hydrolytic stability or corrosion resistance of the fibers (Ehrenstein et

al., 1990). The amount of B2O3 in glass fibers influences the hydrolytic stability of

the glass fiber surface. B2O3, which lowers surface energy of the molten glass, may

concentrate in the surface layers of the glass fiber during the production process

(Loewenstein, 1966). Because B2O3 is exceedingly reactive with water (Loewenstein,

1966), susceptibility of the glass fiber-polymer interface to hydrolytic degradation in

the presence of B2O3 may be increased if the composites are used in an aqueous

environment.

Unidirectional glass fibers are fiber rovings or yarns consisting of 1.000 to 200.000

single glass fibers. Unidirectional fibers give anisotropic mechanical properties to the

composite, and are suitable for applications in which the direction of highest stress is

known. For example, the fibers can be used to efficiently reinforce polymers in one

direction if there is adequate adhesion between the polymer and the fibers. Possible

uses of this type of material in prostheses would include the pontics of the fixed

partial denture and some removable partial denture designs (Vallittu, 1997).

Bidirectional weaves reinforce structures in two directions and are therefore useful

when the direction of highest stress in the prosthesis is difficult to predict. The

rovings or yarns can be woven in different weaves in which the fibers are oriented in

two directions. The weaves may have various textile structures, such as linen, twill,

or satin weave (Airasmaa et al., 1994). The efficiency of bidirectional fiber

16

reinforcement (fibers at a 45° angle to the force) is ½, in the contrast to unidirectional

fiber reinforcement, which has a reinforcing efficiency of 1. Examples of indications

for bidirectional weaves may be resin crowns and some types of removable partial

dentures, such as overdentures.

Table 1. Composition of different glass fibers (wt%)

Components E-glass Acid-

resistant

glass

Alkaline-

resistant

glass

R-glass S-glass

SiO2 53-55 56-58 62 60 62-65

Al2O3 14-16 12 0.8 25 20-25

CaO 20-24 17-22 5.6 6-9 -

MgO * 2-5 - 6-9 -

B2O3 6-9 - - - 0-1

K2O ≤1 0.4 - 0.1 -

Na2O § 0.1-2 14.8 0.4 0-1

Fe2O3 § 0.2-2 - 0.3 0.2

ZrO2 - 2 - - -

ZnO 0-0.7 2 0 - -

* Total amount of CaO and MgO is 20-24wt%-

§ Total amount of K2O, Na2O, and Fe2O3 < 1wt%.

2.3 Impregnation of the fibers

Many authors have investigated the impregnation of the fibers with the matrix

because poor impregnation creates problems using FRC in dentistry. (Miettinen et

al., 1999; Vallittu et al 1998). Fiber reinforcement is only successful if the loading

force can be transferred from the matrix to the fiber. In the case of voids between the

matrix and the fiber, the load-bearing capacity of the FRC decreases.

Poorly impregnated fibers cause another problem: the increase in water absorption in

FRC Issac, 1999; Miettinen et al., 1999; Jancar et al., 1993) which reduces the

mechanical properties (Söderholm et al., 1990; Söderholm et al., 1984). Voids and

cracks in the laminate allows water to enter. A reliable adhesion between the fibers

and the matrix reduces voids and cracks, which can limit the water absorption. In the

case of glass fiber-reinforced reconstructions, the fibers are covered with a silane-

17

coupling agent. Plueddemann postulates a condensation reaction between the silanol

groups and the glass surface (Plueddemann, 1982). The more siloxane bridges that

exist, the less water will be absorbed resulting in more adhesion between matrix and

fibers. Furthermore, the composition of the glass fibers is particularly important. The

content of alkali and earth-alkali ions and boron oxide has to be considered due to the

increased reactivity of these ions and oxides to water. The mechanism of hydrolytic

degradation is based on the leaching effect of boron oxide from the glass surface

(Vallittu, 1999).

It should be noted that by correct treatment of the glass fibers in the sizing procedure,

the corrosion of the glass fiber surface could be diminished. Thus, glass fibers from

different manufactures and with different surface chemistry might behave differently

in this respect. Furthermore, voids of poorly impregnated fibers are oxygen reserves

(Vallittu, 1999). The oxygen inhibits radical polymerization of the polymer matrix.

This decreases the strength of the FRC and increases the residual monomer content,

which can lead to irritant reactions in the oral mucosa (Hensten-Pettersen, 1998)

To solve all these problems, pre-impregnated (pre-pregs) FRC are used. Pre-

impregnated means that the glass fibers are covered with a silane coupling agent and

then pulled through convoluted paths around supports with a bath of light- and/or

heat-curable monomer systems of polymers (Goldberg, 1999).

Pre-pregs of various sizes can be produced to facilitate clinical application. In

dentistry, generally three systems are used to form fiber-reinforced frameworks for

fixed partial dentures with pre-impregnated glass fibers. One system, like Vectris

(Ivoclar, Schaan, FL) is based on a vacuum/pressure adaptation of the fibers in a

mold (Unterbrink, 1999). The purpose of this procedure is to maximize the fiber

content, decrease the number of voids in the framework, and reduce the technique

sensitivity in order to improve the mechanical properties. Other systems, like

FibreKor (Jeneric/Pentron, Wallingford,CT), prefer a manual adaptation of the pre-

impregnated fibers. (Freilich et al., 1998). The advantage of this procedure is said to

reduce the equipment needed in its manufacturing. The third system (Stick, Stick

Tech Ltd, Turku, Finland) is based on pre-impregnated glass fibers with

thermoplastic polymers, which form a multiphase polymer matrix for FRC with

light-curing monomers.

18

2.4 Quantity of fibers

The strength of a fibre-reinforced material depends on the volume content of the

fibres. The better the densification of the glass-fibre, the higher the mechanical

strength will be (Agarwal et al.,1990; Zanghellini, 1997). Highly densified fibre

elements, however, are too rigid to be formed at will. Therefore, Vectris is pressed

into the desired form and simultaneously densified during the forming procedure in

the VS 1. Subsequently, the matrix is polymerized with light, which secures the

shape of the framework (Fig. 1).

Fig.1. Scanning electron microscopic image of the cross-section of Vectris Pontic fibres before (left) and after (right) densification in the Vectris VS1 device. (Courtesy of Dr. Urs Lendenmann)

2.5 Direction of the fibers

In dental reconstructions, unidirectional and bi- or multidirectional fiber orientation

is used. Unidirectional fibers produce anisotropic mechanical properties in the

composite (Goldberg et al., 1994; Issac, 1998; Jauss, 1997; Vallittu, 1998) and are

preferred when the direction of the highest stress is known. In other cases the rovings

can be woven in such a way that the fibers are oriented in two or three directions,

giving the FRC so-called orthotropic mechanical properties (Vallittu, 1998).

However, the efficiency of woven multidirectional fiber reinforcement is reduced as

described in the Krenchel-formula (Elias, 1992). Numerous articles demonstrate the

relationship between the quantity of fibers in the polymer matrix and the

enhancement of the flexural, transverse and impact strength of fiber-reinforced

reconstructions (Zanghellini, 1992) has been described that with increasing fiber

content, the flexural strength increases linearly. The fiber quantity in the polymer

matrix should be defined in volume not weight percentage.

In the case of carbon-, aramid- or ultra-high-modulus polyethylene fibers (UHMPE),

which have a lower density than glass fibers, the fiber content can lead to misleading

results with regard to the strength of FRC (Vallittu, 1998). Vallittu describes a

formula to transform fiber weight percentage into volume percentage (Vallittu,

1997).

19

2.6 Position of fibers

Previous dental FRC research on position and orientation has focused upon the

effects of the question of fiber reinforcement directionality (i.e. random or

longitudinal orientations) (De Boer, 1984; Galan et al., 1989). It is widely accepted

that directional orientation of the fiber long axis perpendicular to an applied force

will result in strength reinforcement. Forces that are parallel to the long axis of the

fibers, however, produce matrix-dominated failures and consequently yield little

actual reinforcement. Design strategies are on occasion employed to provide multi-

directional reinforcement, to minimize the highly anisotropic behavior of

unidirectional fiber reinforcement. Multidirectional reinforcement, however, is

accompanied by a decrease in strength in any one direction when compared with

unidirectional fiber, as described by Krenchel (Vishu, 1998). In most instances in the

dental literature, fiber reinforcement has been positioned in the center of a composite

specimen (De Boer, 1984). Yet from engineering applications, it is known that the

position and orientation of the reinforcement within a construction influences

mechanical properties (Hull, 1990). For a small sized construction, such as a dental

prosthesis, the quality and characteristics of the FRC are important and demand

careful attention. Fiber reinforcement should be optimal when designing prostheses

and their components. As an example, the components (e.g. connector, pontic,

retainer) of a FRC fixed partial denture (FPD) need to be designed to withstand

masticatory loading (Dyer, 2002). While it is known that tension side fiber

reinforcement strengthens a loaded construction, the effect of varying the cross-

sectional design in a FRC structure is not fully known. Respectively, all factors

relating to design and failure of FRC structures should be investigated and better

understood. In conclusion, position and fiber orientation influenced the load to initial

and final failure, and specimen deflection. Tension side reinforcement was most

effective in increasing the load to initial and final fracture (Dyer et al., 2004).

2.7 Water sorption of FRC matrix

Glass fibers are those most often used for reinforcing polymers in prosthetic dentistry

because of the good aesthetic qualities of glass fiber (Vallittu, 1997) and goog

bonding of glass fibers to polymers via silane coupling agents (Rosen, 1978; Mittal,

20

1992; Vallittu, 1997). The most common type of glass used in fiber production is the

so-called E-glass (electrical glass), and this type of glass is also most often used in

dental fiber composites (Vallittu, 1998).

An aqueous environment, such as in the oral cavity, can induce “corrosion” effects in

the surface of glass fibers resulting from water that diffuses through the polymer

matrix (Ehrenstein GW, 1990). This can lead to a reduction of the mechanical

properties and changes in the composite structure, because the surface of the glass

fibers is affected by the hydrolysis of alkali and earth alkali oxides in the glass and

leaching of ions. The composition of the glass is therefore decisive for the hydrolytic

stability of the glass fibers. The silanization used to bond the fibers to the polymer

matrix also influences the hydrolytic stability of the composite (Pantano et al., 1992).

The polymers used in prosthetic dentistry are often multiphase acrylic resin systems

made from prepolymerized powder beads (predominantly poly[methyl methacrylate]

or PMMA) and a liquid of monomers such as methyl methacrylate (MMA) with

ethyleneglycol dimethacrylate (EGDMA) or 1,4-butanediol dimethacrylate (1,4-

BDMA) as cross-linking agents. (Ruyter et al., 1982 ; Öysaed et al., 1982; Öysaed et

al., 1989; Hill, 1981). Water sorption of such multiphase acrylic resins is

approximatively 2wt% (Al-Mulla et al., 1989; Kalanchandra et al., 1987a;

Kalanchandra et al., 1987b; Miettinen et al., 1997). The cross-linking agent EGDMA

has little effect on the water sorption of denture base polymers (Jagger et al., 1990;

Arima et al., 1996). In a fiber-polymer composite, the water sorption is also affected

by the impregnation of fibers with a resin. If there are regions in which the fibers are

not completely embedded with resin, there will be voids in the structure of cured

composite that increase water sorption (Peltonen, 1992; Vallittu, 1995a; Vallittu,

1995b). In conclusion, water has a plasticizing effect resulting from interaction with

the polymer structure (Ruyter et al., 1986). Many studies on the water sorption of

denture base polymers have been carried out, and it has been concluded that water

sorption decreases the mechanical properties of denture base polymers (Hargreaves,

1979).

21

Fig.1

The picture shows the structure of the fiber-

reinforced composite. On the left side there

is the veneering composite (Targis; Ivoclar-

Vivadent) and on the right side the fiber of

pontic (Vectris; Ivoclar-Vivadent) embedded

of Bis-GMA 24.5% and triethylene glycol

dimethacrylates 6.2% with 65% of glass

fibers.

Fig.2

The weave glass fibers of frame (Vectris;

Ivoclar-Vivadent) are moistly constituted of

Bis-GMA 35.2% and triethylene glycol

dimethacrylates 8.8% with 50% of glass

fibers.

Fig.3

The glass fiber is 10 mm of diameter.

References

Agarwal BD, Broutman LJ (1990). Analysis and performance of fiber

composites. 2 ed. New York: John Wiley & Sons.

Airasmaa I, Johansson CJ, Kokko J. Lujitemuovitekniikka (ed 1). Hameenlinna,

Arvi A Karisto Oy 1984 pp 246-271. [in Finnish].

Barbero EJ. Introduction to composite material design. Ann Arbor, MI: Taylor

and Francis; 1998. pp. 2 and 9.

Behr M, Rosentritt M, Lang R, Handel G. Flexural properties of fiber

reinforced composite using a vacuum/pressure or a manual adaptation

manufacturing process. J Dent 2000; 28(7):509—14.

22

Behr M, Rosentritt M, Leibrock A, Schneider-Feyrer S, Handel G. In-vitro

study of fracture strength and marginal adaptation of fibre-reinforced adhesive fixed

partial inlay dentures. J Dent 1999; 27: 163–168.

Chung KH, Ling T, Wang F. Flexural strength of a provisional resin material

with fibre addition. J Oral Rehabil 1998;25: 214—7.

DeBoer J, Vermilyea SG, Brady RE. The effect of carbon fiber orientation on

the fatigue resistance and bending properties of two denture resins. J Prosthet Dent

1984; 51:119—21.

Dyer SR, Lassila LVJ, Jokinen M, Vallittu PK. Effect of fiber position and

orientation load of fiber-reinforced composite. Dent Mater 2004;20:947–955.

Dyer SR. Current design factors in fiber reinforced composite fixed partial

dentures. In: Vallittu PK, editor. The Second International Symposium on Fibre-

Reinforced Plastics in Dentistry. Symposium Book on the Scientific Workshop on

Dental Fibre-Reinforced Composite on 13 October 2001 in Nijmegen, The

Netherlands; 2002.

Ehrenstein GW, Schmiemann A, Bledzki A, Spaude R. Corrosion phenomena

in glass-fiber reinforced thermosetting resins. In:Cheremisinoff NP (ed). Handbook

of ceramics and composites, vol 1. New York: Marcel Dekker, 1990:231-268.

Ehrenstein GW, Schmiemann A, Bledzki A. Corrosion phenomena in glass-

fiber-reinforced thermosetting resins, in Cheremisinoff NP (ed): Handbook of

ceramic and composites (ed 1). New York, Dekker, 1990 pp 231-268.

Ellakwa AE, Shortall AC, Shehata MK, Marquis PM. The influence of fibre

placement and position on the efficiency of reinforcement of fibre reinforced

composite bridgework. J Oral Rehabil 2001;28(8):785—91.

Galan D, Lynch E. The effect of reinforcing fibres in denture acrylics. J Irish

Dent Assoc 1989;35:109—13.

Göhring TN, Peters OA, Lutz F. Marginal adaptation of bonded slot-inlays

anchoring four-unit fixed partial dentures. J Prosthet Dent 2001; 86: 81–92.

Goldberg AJ, Burstone CJ. The use of continuous fiber reinforcement in

dentistry. Dent Mater 1992;8(3):197—202.

Grotsch G Glasfaserverstärkte Kunststoffe - Allgemeines und zahnärztliche

Probleme um glasfaservertärkte Kunststoffe (GFK). Quintessenz 1965b;16:47-9.

23

Grotsch G. Glasfaserverstärkte Kunststoffe (GFK). Quintessenz 1965a;16:109-

10.

Hargreaves AS. Equilibrium water uptake and denture base resin behaviour. J

Dent 1979; 6:342-349).

Hlavác J. Glass technology, Hlavác J (ed): The technology of glass and

ceramic. An Introduction (ed 1). Amsterdam, Elsevier, 1983 pp 55-220.

Hull D. An introduction to composite materials. Cambridge: University Press;

1990. pp. VII, 24—5, 36—7.

Imai T, Yamagata S, Watari, F, Kobayashi, M, Nagayama K, Toyoizumi H,

Uga M, Nakamura S. Temperature-dependence of the mechanical properties of FRP

orthodontic wire. Dent Mater J 1999;18(2):167—75.

Kilfoil BM, Hesby RA, Pelleu Jr GB. The tensile strength of a composite resin

reinforced with carbon fibers. J Prosthet Dent 1983;50(1):40—3.

Körber KH, Körber S. Mechanische Festigkeit von Faserverbundbrücken

Targis/Vectris. ZWR 1996; 105: 693–702.

Ladizesky NH, Ho CF, Chow TW. Reinforcement of complete denture bases

with continuous high performance polyethylene fibers. J Prosthet Dent

1992;68(6):934—9.

Lassila LV, Nohrstrom T, Vallittu PK. The influence of short-term water

storage on the flexural properties of unidirectional glass fiber-reinforced

composites. Biomaterials 2002;23(10):2221—9.

Loewenstein KL. Glass systems in Holliday L (ed). Composite materials (ed 1).

Amsterdam, Elsevier, 1966 pp 129-287.

Malquarti G, Berruet RG, Bois D. Prosthetic use of carbon fiber-reinforced

epoxy resin for esthetic crowns and fixed partial dentures. J Prosthet Dent

1990;63(3):251—7.

Mc Creight LR. Overview of fiber composites. J Dent Res 1967;46:1192.

Meiers JC, Freilich MA. Conservative anterior tooth replacement using fiber-

reinforced composite. Oper Dent 2000; 25(3):239—43.

Mittal KL. Reminiscing of silane coupling agents. In: Mittal (ed). Silanes and

other coupling agents. Utrecht VSP, 1992:3-12.

Nohrstrom TJ, Vallittu PK, Yli-Urpo A. The effect of placement and quantity

24

of glass fibers on the fracture resistance of interim fixed partial dentures. Int J

Prosthodont 2000;13(1):72—8.

Oysaed H, Ruyter IE.Water sorption and filler characteristics of composites for use in posterior teeth. J Dent Res. 1986;65:1315-8.

Pantano CG, Carman LA, Warner S. Glass fiber surface effects in silane

coupling. In: Mittal KL (Ed). Silanes and Other Coupling Agents. Utrecht VSP,

1992:229-240.

Peltonen P Järvelä P. Methodology for determining the degree of impregnation

from continuous glass fibre prepreg. Polymer Testing 1992;11:215-244).

PK, Lassila VP, Lappalainen R. Transverse strength and fatigue of denture

acrylic—glass fiber composite. Dent Mater 1994;10(2):116—21.

Rosen MR. From treating solution to filler surface and beyond. The life history

of a silane-coupling agent. J Coat Tech 1978;50:70-82.

Ruyter IE, Ekstrand K, Bjork N. Development of carbon/graphite fiber

reinforced poly (methyl methacrylate) suitable for implant-fixed dental bridges.

Dent Mater 1986;2(1):6—9.

Ruyter IE, Oysaed H. Conversion in different depths of ultraviolet and visible

light activated composite materials. Acta Odontol Scand. 1982;40:179-92.

Vallittu PK, Lassila VP, Lappalainen R. Acrylic resin-fiber composite--Part I:

The effect of fiber concentration on fracture resistance. J Prosthet Dent

1994;71:607-612.

Vallittu PK. A review of fiber-reinforced denture base resins. J Prosthodontics

1996;5:270-276.

Vallittu PK. Compositional and weave pattern analyses of glass fibers in dental

polymer fiber composites. J Prosthodont 1998;7(3):170—6.

Vallittu PK. Curing of a silane coupling agent and its effect on the transverse

strength of autopolymerizing polymethylmethacrylate-glass fibre composite. J Oral

Rehabil. 1997 Feb;24(2):124-30.

Vallittu PK. Effect of 180-week water storage on the flexural properties of E-

glass and silica fiber acrylic resin composite. Int J Prosthodont 2000;13(4):334—9.

Vallittu PK. Glass fiber reinforcement in repaired acrylic resin removable

dentures: preliminary results of a clinical study. Quintessence Int. 1997

Jan;28(1):39-44.

25

Vallittu PK. Some aspects of the tensile strength of undirectional glass fibre-

polymethyl methacrylate composite used in dentures. J Oral Rehabil. 1998

Feb;25(2):100-5.

Vallittu PK. Strength and interfacial adhesion of FRC-tooth system. In: Vallittu

PK, editor. The Second International Symposium on Fibre-Reinforced Plastics in

Dentistry. Symposium Book on the Scientific Workshop on Dental Fibre-

Reinforced Composite on 13 October 2001 in Nijmegen, The Netherlands; 2002.

Viguie G, Malquarti G, Vincent B, Bourgeois D. Epoxy/carbon composite

resins in dentistry: mechanical properties related to fiber reinforcements. J Prosthet

Dent 1994; 72(3):245—9.

Vishu S. Handbook of plastic testing technology, 2nd ed. New York: John

Wiley; 1998. pp. 546.

Zanghellini G (1997). Faserverstärkung – die Festigkeit ist eine Funktion des

Volumenanteils der Fasern im FRCWerkstoff. Phillip J 14:390-393.

26

Chapter 3: Marginal adaptation of inlay-fixed partial dentures

Missing single-tooth situations offer several reconstructive treatments modalities.

The traditional way is the reconstruction with a conventional metal-ceramic fixed

partial denture (FPD) (Valderhaug 1991). This technique requires a full-coverage

preparation of the abutment teeth. Consequently, a large quantity of sound tooth

structure is destroyed during the preparation (Edelhoff et al 2002). This is

particularly problematic in healthy and young teeth with large pulpal chambers. In

order to limit this destruction and thanks to the evolution of adhesive dentistry,

(Perdigão et al 1999) and implantology, adhesive fixed partial dentures (AFPD)

(Freilich et al 1998) and dental implants ( Leal et al 2001) represent the current

alternatives. These treatments have several advantages over conventional bridges,

especially in relation to conservation of tooth structure and their reversibility (Lutz et

al 2000). Nevertheless when an implant is contraindicated or refused by the patient,

metal-free restorative options may become attractive. Better bonding properties to

composite cements, more appropriate biomechanical behaviour, and enhanced

aesthetics are expected with the use of composite or ceramic compared to metal

alloys. Inlay-, onlay- and partial crown-anchored FPDs can be bonded to the adjacent

teeth and show acceptable short-term results (Göhring et al 2002, Monaco et al

2003). Fiber reinforced composites (FRC) (Krejci et al 1998), high-strength

reinforced ceramics (Edelhoff et al 2001) and a combination of these two materials

(Rosentritt et al 2003) have been proposed for the fabrication of metal-free inlay

fixed partial dentures (IFPDs). Physical data on reinforced composites suggest that

these materials are best suited for conservative inlay FPDs (Göhring et al 1999).

With a carefully executed bonding technique, good results in marginal adaptation

have been achieved with composite inlays (Lutz et al 1991). The stress-resistant

marginal integrity of composite inlays has been attributed to their dentin-like

elasticity modulus (Krejci et al 1994, Braem et al 1995)

The following study regards the different marginal adaptation measured as

continuous/noncontinuous margin of three materials with different Young modulus.

27

FRC are a new material group with a significantly shorter history of use than more

traditional materials. Glass fibers have been reported to considerably improve the

strength of dental polymers when the fibers were silanated and preimpregnated with

the polymer (Karmaker et al 1997). The combination between resin composite and

fiber seems to better comply with stress and provides a straightforward approach in

the laboratory procedure because casting is not necessary (Vallittu 1999). After

simulation of oral stresses, the fracture resistance and marginal adaptation of IFPDs

made with FRC were better than the ones of all-ceramic restorations (Loose et at

1998).

The interest of clinicians in all ceramic systems is rapidly increasing as stronger and

tougher materials are developed and commercialized along with novel processing

technologies. This development has recently led to the application of zirconia-based

ceramics in dentistry. Moreover, the computer aided design-computer aided

manufacturing (CAD-CAM) is among the most recent advances in dental technology

for direct fabrication of all-ceramic restorations (Wiedhahn et al 2000). The

framework must then be veneered with conventional feldspathic porcelain in order to

achieve the appearance of the natural dentition. Adjustments by grinding may then be

required to improve the fitting of the restoration, and sandblasting of the inner

surface of the restoration is commonly used to enhance the adhesion of the luting

agent to the framework (Kern et al 1998). Yttrium oxide is a stabilizing oxide added

to pure zirconia (Y-TZP) to stabilize it at room temperature and to generate a

multiphase material known as partially stabilized zirconia. The high initial strength

and fracture toughness of Y-TZP results from the physical property of partially

stabilized zirconia. The ability of Y-TZP, so-called ‘‘transformation toughening’’, to

transform from tetragonal crystalline structure to a more voluminous monoclinic

structure that helps to prevent crack propagation, contributes to the strength and

toughness of the ceramic (Ardlin 2002, Williams 1997). In vitro studies of Y-TZP

specimens demonstrated a flexural strength of 900 to 1200 MPa. Y-TZP-based

materials have demonstrated a fracture toughness of 9-10 MPa/ m½, which is almost

double the value demonstrated by alumina-based materials, and almost 3 times the

value demonstrated by lithium disilicate–based materials (Christel et al 1989). An in

vitro study evaluating Y-TZP FPDs under static load demonstrated fracture

resistance of more than 2000 N (Tinschert et al 2001).

28

Points in question are the loading forces that can be withstood as well as the quality

of marginal adaptation that might be reached with FRC and high strength ceramic

systems when used for IFPD restorations. The most relevant mechanical properties to

reduce the clinical failures during loading are flexural strength and fracture toughness

but little information is available on IFPDs. Since mechanical failure is mainly

caused by excessive stresses or deformation that can have a destructive effect on

tooth-restoration interface, a full understanding of stress fields developed in the

dental bridge becomes particularly important. On one hand, some studies with finite

element analysis (Magne et al 2002, Tanimoto et al 2004) suggest that inlay FPDs

made with FRC may be a viable alternative to traditional more invasive FPDs.

Resiliency of the composite may prevent the development of harmful stresses at the

adhesive interface, and reinforcement of the fibers may protect the pontic from

excessive strains, resulting in the restoration’s ability to withstand high functional

loads.

On the other hand, zirconia-ceramic IFPDs exhibited the highest resistance to

fracture when compared to metal-ceramic and glass-ceramic and the failures of the

all-ceramic bridges were always cohesive; located at the connector area that represent

the weakest parts of the bridge (Kılıçarslan et al 2004).

These studies provided insight into a number of biomechanical issues, yet they did

not reveal the marginal adaptation at the tooth/restoration interface during occlusion

and clenching. Although there was positive mechanical behaviour of the new tested

materials, further investigations should be performed on the marginal quality of these

materials.

The aim of this in vitro study was to evaluate the marginal adaptation using

quantitative scanning electron microscope analysis of inlay FPDs made with fiber-

reinforced composite and different ceramic high strength materials after simultaneous

thermal cycling and mechanical loading under the simulation of dentinal fluid that

simulated approximately five years of oral service. The null hypothesis was that there

is no difference in marginal adaptation of the IFPDs before and after fatigue using

materials with different flexural strength and Young’s modulus.

29

MATERIALS AND METHODS

Thirty-six human caries-free molars and premolars of nearly identical size with

completed root growth and stored in a 0.1% thymol solution were selected for this

study. The teeth were randomly and equally divided into 3 groups. The apex of each

root was sealed with an adhesive bonding system and resin composite (Optibond FL,

Kerr; Miris, Coltene, Switzerland) without removal of pulpal tissue and fixed with

the composite on aluminum bases.

Afterwards, the teeth and aluminium bases were immersed in an autopolymerizing

resin (Technovit 4071, Heraeus-Kulzer, Friedrichsdorf, Germany) to an apical depth

of two thirds of the root length to create a strong load-resistant support. Each couple

of teeth (one molar and one premolar) were blocked together with the same

autopolymerizing resin at a distance of 10 mm to each other to prevent movement

during the preparation, impression and luting procedures. In this way, the device

simulated an edentulous space resulting from the loss of one molar. A plastic holding

device with 2 holes was used as a support for the inlay bridges. Two rubber dampers

that were slightly taller than the holes were inserted in the holding device. Eccentric

holes were drilled into the rubber dampers to create a larger distance between the

abutments and to increase the tilting of the abutments toward the space when placed

under load. One holding device with the same distance between the rubber dumpers

was created (Fig.1). To simulate the intrapulpal pressure during the cavity

preparations and the luting procedures, a cylindrical cavity was prepared in each

pulpal chamber 1,5 mm below the amelo-cementum junction. A metal tube with a

diameter of 1.4mm was luted into the cavity with the same adhesive and composite

used to fix them on the bases. Through a connecting silicone tube, the pulpar

chamber was evacuated with a vacuum pump (Vacubrand GmbH, & Co, Wertheim,

Germany), filled with a bubble-free mixture of horse serum (PAA Laboratories

GmbH, Linz, Austria) and phosphate-buffered saline solution (PBS;Oxoid Ltd,

Basingstoke, Hampshire, England) with the aid of a 3-way valve, and finally

connected to a serum infusion bottle. This bottle was placed vertically 34cm above

the specimen to simulate the normal hydrostatic pressure of 25mm Hg within the

tooth until the test was terminated.

30

Tooth preparation Different cavity preparations were made on the teeth to simulate a frequent clinical

situation and to create the space accommodation for the different structure

frameworks. The cavities were prepared with the use of rotating diamond bur (80-

25�m grain size, FG 8113NR, 3113NR; Intensiv SA, Viganello Switzerland; Sirius

180 XL red contra-angle handpiece, Micro-Mega, Bresançon, France) with water-

cooling.

Fig.1. The teeth were blocked together

at a distance of 10 mm to prevent any

movement. The device simulated an

edentulous space resulting from the

loss of one molar. The metal tube luted

into the cavity was filled with a bubble-

free mixture of horse serum and

phosphate-buffered saline solution to

simulate the intra pulpal pressure.

Fig.2-a.

Fig.2-b. Fig.2. The onlay preparation in the molar (a) and inlay cavity in the premolar (b) had mesial

margin in enamel (left) and distal margin in dentin (right). The margins were divided in

different portions to analyze the marginal adaptation in a selective way. A-B: occlusal

enamel, B-C, D-E: approximal enamel, C-D: cervical enamel, F-G, H-I: approximal dentin,

G-H: cervical dentin.

The inlay preparation made in the premolar was a MOD cavity with mesial margin in

dentin, 1 mm below the cementum-enamel junction (CEJ), and distal margin in

enamel, 1 mm above the CEJ. The vestibular-palatal width was 3 mm at the cervical

margin that increased to 4 mm at the upper part of the cavity; the cervical preparation

breadth was 2 mm, similar to the occlusal depth. The onlay preparation made in the

molar was a two cusps partial covering with mesial margin in dentin, 1 mm below

the CEJ and distal margin in enamel, 1 mm above the CEJ. The vestibulo-palatal

31

width was like the premolar preparation and the reduction of the cusps was 2.5 mm,

with 2 mm of occlusal depth in the central fossa (Fig.2a-b).

All dentin surfaces were sealed immediately after the tooth preparation with a 3-step

adhesive system (Optibond FL, Kerr; batch n°25881). Phosphoric acid (Ultraetch,

Ultradent) was applied on dentin for 15 seconds and then rinsed for 30”. The primer

was spread on the dentin for 30” with a microbrush without scrubbing and after, the

bonding was applied to the dentin. After a minimal penetration time of 20 seconds,

the resin was air-thinned and polymerized (Optilux 500, Demetron Inc, Danbury,

Conn.) for 60 seconds. Butt joint cavity finishing lines were finished with a diamond

bur (25mm grain size, N° 3113 NR Intensiv SA) by the use of water-cooling under a

stereomicroscope (Leica MZ6). The polymerised bonding was removed with the

same diamond bur only from the cavity enamel finish lines without touching the

sealed dentin. Impressions were made with Imprint II polyvinyl siloxane (3M ESPE)

with a simultaneous mixing technique according the manufacturer’s instructions.

Provisional restorations were made with Fermit N (Ivoclar-Vivadent) and inserted

without interim cement in analogy to the clinical procedure.

Laboratory manufacturing process.

Eighteen inlay bridges were made using three different materials with unlike flexural

strengths and Young’s moduli (Tab.1). Fiber reinforced composite (SR

Adoro/Vectris; Ivoclar/Vivadent) (Fig.3a-b), zirconium oxide-TZP “tetragonal

zirconia polycrystals” (Cercon, DeguDent; Dentsply) and magnesia-partially

stabilized zirconia (DC-Leolux; DCS Dental, Allschwil) covered with silica-based

ceramics were tested in this study (Fig.3a-b).

Fig.3a

Fig.3b Fig.3a-b Lateral view of the IFPD made with (a) fiber reinforced composite (Adoro/Vectris)

and (b) zirconium oxide-TZP tetragonal zirconia (Cercon).

The FRC system (Group 1) consists of two materials: glass-fibers with different

orientation (Vectris) and a microfilled composite (Adoro) for the veneering of the

fiber framework. The design of the fiberglass framework was first pre-modelled with

a photo-curing resin (Spectra Tray, Ivoclar) to obtain the oval shape and its thickness

checked on to the moulding model. This model was embedded in a transparent

32

silicone impression paste (Transil) to form a mould. Then this resin was removed and

the fibers were applied into the silicone-mould. The pre-impregnated ‘pontic’ fibers

were condensed in a deep-drawing, polymerization process. After a cycle of vacuum-

forming process and then cured by light in VS1 unit (Ivoclar-Vivadent) for 10 min.

according to the manufacturer’s recommendations, the FRC was sandblasted with

Rocatec system (3M ESPE) with small grain size of 80µm at 2,5 bar of pressure for

10 seconds and treated with silane (Wetting agent, Ivoclar-Vivadent). A sheet of

wave fibers ‘frame’ was placed upon the ‘pontic’ structure and the cycle in VS1 was

repeated. The Adoro material was built incrementally using the Quick pre-curing

light unit. The final polymerization / tempering was performed in the Lumamat 100

unit by means of light and heat. The additional tempering step at 104°C was done to

maximize the strength and the surface quality of the restorations.

Cercon (Group 2) is a CAM system that can produce a framework of zirconium

oxide-TZP. The Cercon brain machine automatically mills the framework from an

unsintered zirconium oxide blank (Cercon Base). After that the chalky-soft state is

sintered in the Cercon heat furnace at 1350°C. Finally, the framework is veneered

with low-fusing dental ceramic (Cercon ceram S), which is specially tailored to the

coefficient of thermal expansion of zirconium oxide.

The principle of the Precident system DCS (Group 3) is based on touchless contact-

free measurement and milling in a CAD/CAM process. These two operations are

separated for organizational reasons. The data of the abutments are taken with the

help of a non-contact laser (Preciscan) that at maximum resolution can take 300.000

points/minute. The acquired data are transferred by modem to the milling machine

(Precimill) that can prepare the sub-structure from a sintered magnesia-partially

stabilized zirconia DC-Leolux. Finally the framework is covered with low-fusing

ceramic (Cercon ceram S). The framework of the ceramic bridges (group 2 and 3)

was extended until 1 mm of the margins of the cavity preparation in order to have

etchable silica-based ceramic on the closing margins and optimize the adhesion with

tooth tissue. All the connections of the inlay/onlay with the pontic elements were

3.5X3.5 mm.

Adhesive Procedure

33

The provisional restorations were removed and the inner surfaces of the teeth

previously sealed with bonding were sandblasted with CoJet system (3M ESPE) with

small grain size of 30 µm at 2 bar of pressure for 2 seconds. The inner surfaces of

the FRC and only the zirconium area of the ceramic bridges were treated with CoJet

system (30µm at 2 bar x 10s). The closing ceramic margins were etched with 10%

hydrofluoric acid for 60 seconds and 2 layers of silane-coupling agent (Monobond S,

Ivoclar Vivadent) were applied and heated for 1 minute (ID 500; Coltene

Switzerland) on all the inner surfaces. All enamel and dentin surfaces were luted with

Optibond FL and Tetric Transparent (Ivoclar-Vivadent) by applying the ultrasonic

technique according to the manufacturer’s instructions. The luting cement was light

activated for 60 sec. each from cervical, buccal, lingual and occlusal surfaces. The

margins of the restorations were then finished with 15µm diamond burs

(Composhape, Intensiv) and polished with a composite finishing and polishing kit

(Hawe Neos Dental) in a slow-speed handpiece (Fig.4).

Fig.4 Adhesive inlay

bridge made in fiber

reinforced composite

after the luting

procedures.

Evaluation

The samples were cleaned with rotating nylon brushes (Hawe Neos) and toothpaste

(Signal Anti Caries) before making the impressions for the replicas. Seven partial

impressions for each bridge before and after the thermal and mechanical test were

taken to compare the quality of the marginal adaptation. Six different categories

(approximal enamel, approximal dentin, cervical enamel, cervical dentin, occlusal

and buccal enamel) were recorded to identify the areas with greater stress (Fig. 5).

Gold sputtered (SCD 030, Provac, FL-9496 Balzers, Liechtenstein) epoxy resin

replicas (Epofix, Struers, D-2610 Rodovre, Denmark) of all samples were fabricated

by using polyvinylsiloxane impressions (President Plus Light-body, Colténe AG,

Altstätten, Switzerland). They were subjected to a quantitative evaluation of marginal

34

adaptation at a standard 200x magnification in the SEM (XL20, Philips, NL-5600

Eindhoven, Netherlands) by using a custom made module programmed within an

image processing software (Scion Image, Scion Corp, Frederik, MA 21703, USA).

All specimens were subjected to the quantitative evaluation and examined for

continuous margins (no gap, no interruption of continuity), non-continuous margins

(gap due to adhesive or cohesive failure; fracture of restorative material or fracture of

enamel related to restoration margins), overhangs and underfilled margins. The

percentages of continuous/non-continuous margin were evaluated separately for

tooth-luting composite and luting composite-restoration interfaces. The specimens

were mechanically loaded at the vestibular cusp of the pontic element in a computer-

controlled masticator with 1.200.000 cycles of 49 N each, at a frequency of 1.7 Hz. A

total of 3,000 thermocycles of type 5°C to 55°C to 5°C were performed

simultaneously (Fig.6). The chamber was automatically emptied after 2 minutes for

10 s with air pressure to avoid mixing the cold and warm water (Krejci et al 1990a,

Krejci et al 1993). By having the specimen holders mounted on a rubber rest, a

sliding movement of the bridges was produced during the loading. These conditions

are believed to simulate approximately five years of clinical service (Krejci et al

1990b, Krejci et al 2003). Differences in means were compared with the use of

matched pairs t tests and one-way analysis of variance (ANOVA). The level of

significance was set at P=0.05.

35

36

legend

Kind of replica

P: premolar, M: molar

1. DM

Distal molar

Approximal enamel: a-b; c-d

Cervical enamel: b-c

2. MP

Mesial premolar

Approximal dentin: a-b; c-d Cervical dentin: b-c

3.

VP VM BM

Vestibolar premolar Vestibolar molar Buccal molar

P; Approximal enamel: a-b M; Cervical dentin: c-d Approximal dentin: d-e Buccal enamel: e-f

4. PP

PM

Palatal premolar Palatal molar

P; Approximal enamel: a-b M; Approximal dentin: c-d Cervical dentin: d-e

5. OP OM

Occlusal premolar Occlusal molar

P; occlusal enamel: a-b; c-d M; occlusal enamel: e-f

6. GP

7. GM

Gingival premolar Gingival molar

P; cervical enamel: a-b M; cervical dentin: c-d

Fig. 5. Outline of the non-destructive replica technique.

a

d

cba

b c

d

b

a

e c

d

f

b

a c

e d

a b

c d

e f

a b c d

37

Fig.6. (Left) Loading machine with six watertight cells (A) and the thermocycle device (B).

(Right) The arrow down indicates the rubber dampers that increase the tilting of the

abutments when placed under load. The finger point to the silicone tube filled with a mixture

of horse serum and phosphate-buffered saline solution to simulate the intra pulpal pressure

during all stress cycles. The dot arrow shows the level of the water during the thermocycles.

RESULTS

All restorations were in place after completing the stress test, meaning that the

retention amounted to 100% for all groups. Neither restoration nor abutment

fractures was found after fatigue loading. Only two hairline fractures of the veneering

material that spread in the buccal and vestibular area were found in the gingival part

of the connection between the pontic and the abutment tooth in the FRC group (Fig.

10a-b).

Marginal adaptation was analyzed at the interface of the luting composite and the

abutment inlay/onlay (CI) and at the interface of the tooth and luting composite (TC).

The results of the marginal adaptation expressed in percentage are represented in

table 1. Significant statistical differences (P<0.05) were found for all groups before

and after loading concerning the percentage of continuous margins (CM) as the total

marginal length at the luting cement-restoration and luting cement-tooth interfaces.

No differences were observed after the cycle test between the three groups at the

luting cement-restoration interface (Fig.7). However, significant statistical

differences were found after loading between the FRC and the other two ceramic

systems at the luting cement-tooth interfaces (Fig.8).

The prevailing marginal defect in all groups was pure marginal opening (Fig.11a-b).

Some fractures pointed out after the final observation were traced back as enamel-

dentinal fractures (EF) and filling fracture (FF). No significant difference was

detected in the sub-fracture of the dental tissue (EF) near the margin between the

three groups. However, significant changes (P<0.05) were found in hairline cracks in

the restoration (FF) along the margins between DC-Leolux (4.1%), FRC (0.4) and

Cercon (1.7) after loading. In some cases non continuous “pure” margin identified as

only “open margin” changed in EF or FF. No more than 0.5% of the “overhangs” and

38

“underfilled margins” were found before and after loading, with no significant

differences among the groups. No difference in “continuous margin” was detected

between approximal enamel and approximal dentin. The inner comparison of the

same groups between the onlay preparation (molar) and the inlay cavity (premolar)

didn’t show significant difference (P>0.05). Severe changes in continuous margin

were detected at the tooth-luting composite interface in the dentinal margin after the

test. The values were 20.8% for group 1, 53.8% for group 2 and 32.2% for the last

group Statistical difference was found between Cercon and the other two groups

(P<0.05).

Luting composite-inlay

interface (CI)

Adoro/Vectris Cercon DC Leolux

Before loading 94.6 ± 3.1 92.9 ± 5 96.2 ± 2.1

After loading 88 ± 6.7 85.7 ± 6.1 82.2 ± 9.8

Luting composite-tooth interface (TC)

Before loading 86.7 ± 6.7 93.3 ± 3.4 96.1 ± 2.4

After loading 62.5 ± 16.4 83.2 5.9 75.3 ± 7

Table 1. Percentage of “continuous margin” for the total marginal length before and

after loading (means±SD) at the luting composite-inlay and composite-tooth

interfaces.

39

65

70

75

80

85

90

95

100

Adoro/Vectris before

DC-Leoluxbefore

DC-Leoluxafter

Adoro/Vectris after

Cerconbefore

Cerconafter

Mar

gina

lada

ptat

ion

lutin

g co

mpo

site

-rest

orat

ion

inte

rface

65

70

75

80

85

90

95

100

Adoro/Vectris before

DC-Leoluxbefore

DC-Leoluxafter

Adoro/Vectris after

Cerconbefore

Cerconafter

Mar

gina

lada

ptat

ion

lutin

g co

mpo

site

-rest

orat

ion

inte

rface

Fig. 7. Continuous margin at luting composite-restoration interface (CI) with

quantilies (red line), means/Anova (green lines), means and standard deviation

(blue line).

30

40

50

60

70

80

90

100

1 A BEFORE 1 B AFTER 2 A BEFORE 2 B AFTER 3 A BEFORE 3 B AFTER

Mar

gina

lada

ptat

ion

lutin

g co

mpo

site

-toot

hin

terfa

ce

Adoro/Vectris before

DC-Leoluxbefore

DC-Leoluxafter

Adoro/Vectris after

Cerconbefore

Cerconafter

30

40

50

60

70

80

90

100

1 A BEFORE 1 B AFTER 2 A BEFORE 2 B AFTER 3 A BEFORE 3 B AFTER

Mar

gina

lada

ptat

ion

lutin

g co

mpo

site

-toot

hin

terfa

ce

Adoro/Vectris before

DC-Leoluxbefore

DC-Leoluxafter

Adoro/Vectris after

Cerconbefore

Cerconafter

40

Fig. 8. Continuous margin at tooth-luting composite (TC).

Fig. 9. Continuous margin

of FRC restoration. The left

area (A) represent the

enamel, the middle part the

luting cement and the right

share shows the restoration

(C). The finger points to the

luting cement-tooth

interface whereas the arrow

indicates the luting cement-

restoration interface.

Fig. 10a.

Fig. 10b. Fig. 10a. Hairline fracture of the veneering material in the gingival part of the connection

between the pontic and the abutment in FRC group. The finger indicates the micro crack that

spread in the vestibular area. The black frame in the upper part clearly shows at 200

magnifications (fig.10b) the fissure in the resin composite.

Fig. 11a.

Fig. 11b. Fig. 11. These images show the same portion before (a) and after (b) the stress cycles. The

inner area (A) represent the ceramic restoration (Cercon), the luting cement constitute the

middle part (B), whereas the upper portion (C) show the dental tissue. The arrows and the

fingers indicate the “continuous” (10a) and “non continuous” margins (10b) as result of the 5-

year simulation period.

DISCUSSION

41

Although this study might have some limitations in respect to its clinical relevance,

the absence of detachments or fractures of the inlay bridges suggest that both ceramic

and fiber reinforced composite systems could be utilized in clinical practice.

Nevertheless, some remarks must be made in regard to the quality of the margins and

the hairline fractures found in the FRC group. The most critical areas in dental

bridges and particularly in IFPDs are represented by the connection at the gingival

portion of the pontic between the abutments because this surface constitutes the

tensile side of the beam (Magne et al 1999). When occlusal forces are applied

directly to the long axis of the bridge at the midspan (pontic), compressive stresses

will develop at the occlusal aspect of the connector at the marginal ridge, and tensile

stresses will develop at the gingival surface of the connector (Kelly et al 1995).

These tensile stresses could contribute to the propagation of micro cracks located at

the gingival surface of the connector through the veneering material in an occlusal

direction, and may eventually result in fracture of the composite. The presence of

hairline fractures in the gingival area of the pontic in two bridges of FRC group could

be related to the greater flexibility of the fiber framework compared to the ceramic

materials supported by zirconia framework. These micro cracks can compensate the

smaller stiffness of the fiber but at any rate could determine the beginning of the

delamination or fracture of the layering material.

The clinical fracture resistance of IFPDs is related to the size, shape, and position of

the connectors and to the span of the pontic. The basis for the proper design of the

connectors and the pontic is the law of beams: deflection of a beam increases as the

cube of its length, it is inversely proportional to its width, and it is inversely

proportional to the cube of its height (Raigrodski 2004). Moreover, the flexibility of

the beam is in direct relation to the amount and the type of the fibers that go to make

up the framework. The position of the FRC layer had an effect on the flexural

strength of the test specimen. The highest flexural strength was achieved when the

FRC layer was located at the tension side of the test specimens. The particulate filler

composite is the weakest phase of the test specimen. When it is located on the

tension side, the fracture can easily initiate. The FRC structure benefits most when

the tensile stresses can be transferred to the reinforcing fibers. The veneering

42

particulate filler composite is strong in compression stress and, therefore, the FRC

structure requires less reinforcement fibers on the compression side (Murphy 1998).

Usually, it is preferable to place the FRC laminates symmetrically relating to the

FRC framework, to prevent polymerization shrinkage effect, thermal stresses, and

possible deformation during polymerization (Lassila et a2004).

Nevertheless, it is often very difficult to design the FRC framework to form an

optimal design because of the abutment location and occlusal parameters.

One theoretical assertion is that lower-elastic modulus frameworks would determine

a better stress transfer to the tooth and reduce tensile stresses at the adhesive interface

(Vallittu et al 2000), even though no scientific evidence has shown this to be true.

Vallittu (Vallittu et al 2004) supposed that a lower modulus of elasticity might allow

the FPD to deflect to some extent during function without the formation of stresses

that may cause debonding. Brunton (Brunton et al 1999) preferred restorative

materials such as fiber-reinforced composites rather than ceramic materials because

of the material’s flexibility, repairable properties, and equivalent fracture resistance.

They reported that fiber-reinforced composite material showed similar fracture

resistance when compared to ceramic material under compressive loads for posterior

restorations.

Contrary to these results, in our study the direct comparison between fiber-reinforced

composite and ceramic reinforced systems suggests that different materials could

have an influence on the quality of the margin primarily at the luting composite-tooth

interface. In any case the null hypothesis was rejected. The statistical difference

expressed in percentage between the FRC and the other all-ceramic restorations

could be related to the different flexibility of the frameworks. Our results could

suggest that the smaller stiffness of the fiber-composite complex can negatively

influence the marginal adaptation under load. The fibre framework may absorb the

stress generated during loading but the increased flexibility might have led to the

opening of the margins.

Any significant difference between “approximal enamel” and “approximal dentin”

was found within each group for both interfaces after the fatigue test. All these

margins are in enamel but deferred from the base of the cavity box. The first margin

continues in the “cervical enamel”, the second keeps up in the “cervical dentin”. The

opening of the margin in the cervical dentin doesn’t have an influence on the

43

overhanging enamel. The bonding between the luting composite and the enamel is so

strong that the gap created at the dentin interface stopped at the cementum-enamel

junction. Marginal adaptation at the dentinal margins decreased in a dramatic way

after mechanical loading. The percentage of continuous margin changed from 21 to

54% after the test. A statistical difference was found between the stiffer system

(Cercon) and the other two groups. In any case, the disintegration of the margins in

dentin is so high in all groups that the IFPDs could be contraindicated when one or

both abutments have margins in dentin until the adhesion between the luting

composite and the dentin is improved.

The marginal adaptation at the interface between the luting composite and restoration

decreased after mechanical and thermal loading but no significant difference was

found between all groups. The values ranged between 82,2 and 88%.

Successful ceramic-resin bonding is achieved by the formation of chemical bonds

and micromechanical interlocking at the resin-ceramic interface. With conventional

silica-based ceramics, acid etching and application of a silane coupling agent create a

rough surface of increased wettability for successful ceramic resin bonds. Zirconium-

oxide ceramics are not silica based and the application of acidic agents, such as

hydrofluoric acid, does not create a sufficiently roughened surface for enhanced

micromechanical retention. Advances in adhesive dentistry have resulted in the

recent introduction of modern surface conditioning methods such as silica coating

that require airborne particle abrasion of the surface before bonding in order to

achieve high bond strength. In this technique, the surfaces are air abraded with

aluminum oxide particles modified with silica (Kern et al 1998, Özcan 2002). The

blasting pressure results in the embedding of silica into the ceramic surface,

rendering the silica-modified surface more chemically reactive for the resin through

silane coupling agents. The tribochemical silica coating followed by silanization,

which increased the silica content on the ceramic surface, evidently enhanced the

bond between the ceramic surfaces and the luting cement. Since the silica layer is

well attached to the ceramic surface, this provides a basis for silanes to enhance the

resin bond. Airborn particle abrasion with Al2O3 abrasive particles has proven to be

effective both for composite, aluminum- and zirconium-oxide ceramics (Derand et al

2000). In our study the adhesion between dental tissue and all ceramic bridges was

increased leaving one millimetre or more of silica based ceramic along the margins

44

without zirconia at the interface. This treatment can explain the good results of the

marginal adaptation at both adhesive interfaces of the all-ceramic systems.

Within the limitations of the experimental study, several conclusions can be drawn.

The flexibility of the framework may play an important role in the marginal

adaptation of adhesive inlay/onlay bridges. More rigid materials may transfer the

stress to the margin to smaller degree than flexible materials, which may result in a

more stable bond to the dental tissues under load. When FRC are used for IFPDs

high fiber volume fraction and well-designed framework shape must is mandatory to

increase at maximum the stiffness of the inlay bridges. All ceramic systems

reinforced with zirconia could be used for inlay FPDs in the clinical practice but

simplified CAD/CAM technique is required to allow faster construction of the

zirconia framework. As the marginal adaptation in dentin after load was low in all

groups tested, the IFPDs might be contraindicated when abutments’ margins reach

dentin independent of the bridge material used, until the adhesion between the luting

cement and the dentin will be improved.

45

References

Ardlin B.I. Transformation-toughned zirconia for dental inlays, crowns and

bridges: chemical stability and effect of low-temperature aging on flexural strength

and surface structure. Dental Materials 2002; 18:590-95.

Braem MJ, Lambrechts P, Gladys S, Vanherle G. In vitro fatigue behavior of

restorative composites and glass ionomers. Dent Mater 1995;11:137- 41.

Brunton PA, Cattell P, Burke FJ, Wilson NH. Fracture resistance of teeth

restored with onlays of three contemporary tooth-colored resin-bonded restorative

materials. The Journal of Prosthetic Dentistry 1999; 82:167-71.

Christel P, Meunier A, Heller M, Torre JP, Peille CN. Mechanical properties

and short-term in-vivo evaluation of yttrium-oxide-partially-stabilized zirconia.

Journal of Biomedical Materials Research 1989; 23:45-61.

Derand P, Derand T. Bond strength of luting cements to zirconium oxide

ceramics. The International Journal of Prosthodontics 2000; 13:131-35.

Edelhoff D, Sorensen JA. Tooth structure removal associated with various

preparation designs for posterior teeth. The International Journal of Periodontics &

Restorative Dentistry 2002; 22:241-49.

Edelhoff D, Spiekermann H, Yildirim M. Metal-free inlay-retained fixed partial

dentures. Quintessence International 2001; 32:269-81.

Freilich MA, Duncan JP, Meiers JC. Preimpregnated, fiber-reinforced

prostheses. Part I. Basic rationale and complete-coverage and intracoronal fixed

partial denture designs. Quintessence International 1998; 29:689-96.

Göhring TN, Mörmann WH, Lutz F. Clinical and scanning electron

microscopic evaluation of fiber-reinforced inlay fixed partial dentures: preliminary

results after one year. J Prosthet Dent 1999;82:662-8.

Göhring TN, Schmidlin PR, Lutz F. Two-year clinical and SEM evaluation of

glass-fiber-reinforced inlay fixed partial dentures. American Journal of Dentistry

2002; 15:35-40.

Karmaker AC, Di Benedetto AT, Goldberg AJ. Continuous fiber reinforced

composite materials as alternatives for metal alloys used for dental applicances.

Journal of Biomaterials Applications 1997; 11:318-28.

Kelly JR, Tesk JA, Sorensen JA. Failure of all-ceramic fixed partial dentures in

46

vitro and in vivo: analysis and modeling. Journal of Dental Research 1995; 74:1253-

8.

Kern M, Wegner SM. Bonding to zirconia ceramic: adhesion methods and their

durability. Dental Materials 1998; 14:64-71

Kern M, Wegner SM. Bonding to zirconia ceramic: adhesion methods and their

durability. Dental Materials 1998; 14: 64–71.

Kılıçarslan MA, Kedici PS, Küçükeşmen HC, DDS, Uludag BC. In vitro

fracture resistance of posterior metal-ceramic and all-ceramic inlay-retained resin-

bonded fixed partial dentures. The Journal of Prosthetic Dentistry 2004; 92:365-70.

Krejci I, Boretti R, Giezendanner P, Lutz F. Adhesive crowns and fixed partial

dentures fabricated of ceromer/FRC: clinical and laboratory procedures. Practical

Periodontics and Aesthetic Dentistry 1998; 10:487-98.

Krejci I, Duc O, Dietschi D, de Campos E. Marginal adaptation, retention and

fracture resistance of adhesive composite restorations on devital teeth with and

without posts. Operative Dentistry 2003; 28:127-35.

Krejci I, Kuster M, Lutz F. Influence of dentinal fluid and stress on marginal

adaptation of resin composites. Journal of Dental Research 1993; 72:490-94.

Krejci I, Lutz F, Gautschi L. Wear and marginal adaptation of composite resin

inlays. J Prosthet Dent 1994;72:233-44.

Krejci I, Lutz F. In-vitro test results of the evaluation of dental restoration

systems. Correlation with in-vivo results (in German) Schweizer Monatsschrift fur

Zahnmedizin 1990; 100:1445-49.

Krejci I, Reich T, Lutz F, Albertoni M. An in vitro test procedure for evaluating

dental restoration systems. 1. A computer-controlled mastication simulator (in

German) Schweizer Monatsschrift fur Zahnmedizin 1990; 100:953-60.

Lassila LVJ, Vallittu PK. The effect of fiber position and polymerization

condition on the flexural properties of fiber-reinforced composite. The Journal of

Contemporary Dental Practice 2004; 2:014-026.

Lassila LVJ, Vallittu PK. The effect of fiber position and polymerization

condition on the flexural properties of fiber-reinforced composite. The Journal of

Contemporary Dental Practice 2004; 2:014-026.

Leal FR, Cobb DS, Denehy GE, Margeas RC. A conservative aesthetic solution

47

for a single anterior edentulous space: case report and one-year follow-up. Practical

Procedures & Aesthetic Dentistry 2001; 13:635-41.

Loose M, Rosentritt M, Leibrock A, Behr M, Handel G. In vitro study of

fracture strength and marginal adaptation of fiber-reinforced-composite versus all

ceramic fixed partial dentures. The European Journal of Prosthodontics and

Restorative Dentistry 1998; 6:55-62.

Lutz F, Göhring TN. Fiber-reinforced inlay fixed partial dentures: maximum

preservation of dental hard tissue. Journal of Esthetic Dentistry 2000; 3:164-71.

Lutz F, Krejci I, Barbakow f. Quality and durability of marginal adaptation in

bonded composite restorations. Dent Mater 1991;7:107-13.

Magne P, Perakis N, Belser U, Krejci I. Stress distribution of inlay-anchored

adhesive fixed partial dentures: a finite element analysis of the influence of

restorative materials and abutment preparation design. The Journal of Prosthetic

Dentistry 2002; 87:516-27.

Magne P, Versluis A, Douglas WH. Rationalization of incisor shape:

experimental-numerical analysis. The Journal of Prosthetic Dentistry 1999; 81:345-

55.

Monaco C, Ferrari M, Miceli GP, Scotti R. Clinical evaluation of fiber-

reinforced composite inlay FPDs. The International Journal of Prosthodontics 2003;

16:319-25.

Murphy J. Reinforced plastics handbook. 2nd ed. Oxford: Elsevier Science;

1998.p. 237-67.

Özcan M. The use of chairside silica coating for different dental applications: a

clinical report. The Journal of Prosthetic Dentistry 2002; 87:469–72.

Perdigão J, Lopes M. Dentin bonding-state of art 1999. The Compendium of

Continuing Education in Dentistry 1999; 20:1151-62.

Raigrodski A. Contemporary materials and technologies for all-ceramic fixed

partial dentures: A review of the literature. The Journal of Prosthetic Dentistry 2004;

92:557-562.

Rosentritt M, Behr M, Handel G. Fixed partial dentures: all-ceramics, fibre-

reinforced composites and experimental systems. Journal of Oral Rehabilitation

2003; 30:873-77.

48

Tanimoto Y, Nishiwaki T, Nemoto K. Numerical failure of glass-fiber-

reinforced composites. Journal of Biomedical Materials Research. 2004; 68A:107-

13.

Tinschert J, Natt G, Mautsch W, Augthun M, Spiekermann H. Fracture

resistance of lithium disilicate-, alumina-, and zirconia-based three-unit fixed partial

dentures: a laboratory study. The International Journal of Prosthodontics 2001;

14:231-38.

Valderhaug J. A 15-year clinical evaluation of fixed prosthodontics. Acta

Odontologica Scandinavica 1991;49:35-40.

Vallittu PK, Sevelius C. Resin-bonded, glass fiber-reinforced composite fixed

partial dentures: a clinical study. The Journal of Prosthetic Dentistry 2000; 84:413-

418.

Vallittu PK. Prosthodontic treatment with a glass fiber-reinforced resin-bonded

fixed partial denture: a clinical report. The Journal of Prosthetic Dentistry 1999;

82:132-35

Vallittu PK. Survival rates of resin-bonded, glass fiber–reinforced composite

fixed partial dentures with a mean follow-up of 42 months: A pilot study. The

Journal of Prosthetic Dentistry 2004; 91:241-246.

Wiedhahn K. Cerec 3 as the first step in digitizing the dental office.

International Journal of Computerized Dentistry 2000; 3:67-71.

Williams D. Ceramics transformed: manipulating crystal structures to toughen

bioceramics. Medical Device Technology 1997; 8:6-8.

49

Chapter 4: Criteria for selecting the materials for IFPDs

50

In case of single missing posterior teeth, the adjacent teeth often contain small

occlusal or proximal restorations or initial carious lesions. Conventional porcelain

fused to metal fixed partial dentures (FPDs) provide high fracture strength and have a

long clinical history (Freilich et al 1998, Pauli 1996). But in most cases full or partial

coverage preparations are needed to stabilize the remaining tooth structure and

provide sufficient retention. A significant amount of sound tooth structure must be

removed when fabricating conventional FPDs (Rammelsberg et al 1995).

Fiber-reinforced composites are a new material group with a significantly shorter

history of clinical use than more traditional prosthodontic materials. These materials

differ in the preparation of their laminates (Behr et al 2000, Behr et al 2001).

In posterior areas, however, high mastication forces (approximately 500 to 600 N10)

may restrict the use of metal-free IFPDs. The aim of this in vitro study was to

examine the fracture strength of metal-free 3-unit posterior IFPDs after aging in an

artificial oral environment. One glass fiber–reinforced composites and 2 all-ceramic

systems were tested.

Materials and Methods

The previously eighteen inlay bridges made using three different materials with

unlike flexural strengths and Young’s moduli were loaded until the fracture after the

fatigue machine. FRC (SR Adoro/Vectris; Ivoclar/Vivadent), zirconium oxide-TZP

“tetragonal zirconia polycrystals” (Cercon, DeguDent; Dentsply) and magnesia-

partially stabilized zirconia (DC-Leolux; DCS Dental, Allschwil) covered with silica-

based ceramics were tested in this study. The fracture strength of each IFPD was

determined by mechanically loading them to failure with a universal testing machine

(Instron) after the aforementioned artificial aging. The force was applied on the

center of the pontics using a steel ball (6 mm diameter, 1 mm/minute crosshead

speed). To ensure regular force distribution and minimize the transmission of local

force peaks from the steel ball to the cusps of the pontics, a layer of tin foil (0.5 mm)

was inserted. The failure determination was set at a 10% loss of the maximum

loading force. Radiologic examinations (X-ray, Castellini, Italy) were made to

document the different fracture patterns. Median standard deviation values were

calculated for all groups. The statistical analysis was performed using the Mann-

51

Whitney U test and the Kruskal- Wallis test for nonnormal distributed values (JMP

5.1). The level of significance was set to p =0.05.

Results

All fracture strength results are shown in Figures 1-2. The results for all samples are

given in Table 1. The values for FRC Adoro/Vectris system was 1382N (1172N 25%

percentile, 1549N 75% percentile), for DSC DC-Leolux 1433N (1239N 25%

percentile, 1570N 75% percentile), and for Cercon system was 1695N (1538N 25%

percentile, 1814N 75% percentile). There was no statistical difference between FRC

and DC-Leolux, and DC-Leolux and the Cercon system. In contrast, there was a

statistical difference between the FRC group and the Cercon system. Fig. 3 shows a

representative pattern of fracture for a fiber-reinforced composite IFPD. The failure

of the IFPDs can be attributed to either fractures of the facing material or shearing of

the facing material from the framework fibers. No discernible damage to the

framework fibers was observed. Fig 4 shows the characteristic failure pattern of an

allceramic IFPD. The fractures occurred mainly at the connectors between the pontic

and the abutment in combination with a complete fracture of the restoration.

Table. Results for all samples FRC DC-Leolux Cercon 1 1613 1789 1846 2 1528 1497 1804 3 1467 1440 1744 4 1297 1426 1646 5 1177 1278 1580 6 1160 1123 1413 Total 1373,7 1425,5 1672,2 Standard deviation 189,7 223,9 160,7

52

a

0

500

1000

1500

2000

FRC

FRC 1613 1528 1467,7 1297 1177 1160

1 2 3 4 5 6

b

0

200

400

600

800

1000

1200

1400

1600

1800

DCS

DCS 1789 1497 1440 1426 1278 1123

1 2 3 4 5 6

Fig. 1-a-b. The graphs a, b showes the values of the FRC and DSC groups.

a

0

500

1000

1500

2000

CERCON

CERCON 1846 1804 1744 1646 1580 1413

1 2 3 4 5 6

Fig. 2. The graph a shows the value of the Cercon group for all samples.

Fig. 3. The graph a shows the comparition of the all groups.

Col

umn

1

1100

1200

1300

1400

1500

1600

1700

1800

1900

Cercon DC-Leolux FRC

53

Discussion

The examination methods used for testing fracture strength follow established

procedures previously described in the literature ( Pauli 1996, Behr et al 2000, Behr

et al 2001). The artificial oral environment allowed for the simulation of parameters

such as temperature changes, mastication force, frequency of occlusal loads, and

resiliency of the human periodontium (Krejci et al 1990). The advantage of this

system is the application of similar conditions on each specimen. This system can be

used for aging dental restorations in vitro to estimate mechanical properties of new

dental materials before any timeconsuming and large-scale clinical investigations are

undertaken (Rosentritt et al 1997, Delong et al 1983). Assuming maximum

mastication forces of about 500 N (Hidaka et al 1999) in posterior chewing areas, the

results found in this study were within the range for the fiber-reinforced systems

(Adoro/Vectris) and for the all-ceramic system (DC-Leolux, Cercon).

The results for the Adoro/Vectris IFPD are increased to those found in a vitro study

conducted by Behr (Behr et al 1999) comparing tub-shaped and box-shaped

preparation techniques with Targis/Vectris. In this study the framework were

composed only of the pontic prepreg fibers and showed fracture strength values of

696 N (531 N/958 N) in the box-shaped preparation and 722 N (665 N/818 N) in the

tub-shaped preparation. The Vectris framework consists of 3 different prepregs that

can be distinguished by the fiber orientation; prepregs with parallel fibers are called

“pontic,” those with a 45-degree alignment are called “single,” and those with a 90-

degree alignment are the “frame” prepregs. Kolbeck (Kolbeck et al 2002)

investigated with the Targis/vectris system a combination of pontic prepreg and

single prepreg fibers and showed nearly exactly the same fracture strength than those

without the single prepreg [723 N (692N/806 N)]. Thus, one might interpret that for

Vectris/Targis IFPDs, it is more important to prepare a tub-shaped cavity without

sharp inner edges rather than to use a large quantity of different prepreg fibers. A

disadvantage of the tub-shaped cavity might be the greater potential for pulp

irritation or exposure. In any cases, the value for the FRC group was lower than

1000N and rather near to the threshold of the 500N. The laboratory technique used in

our study allows the extension in the vestibular and buccal side to support the

layering material like in metal framework and hold it in all loading directions. The

framework in the FRC was made in an anatomical shape in the pontic element with

54

parallel and weave fiberglasses. Vestibular and buccal FRC extensions in the pontic

element can increase the bonding area between the framework and resin composite as

well as hold the veneer material better during occlusal loading. This observation

suggested that modified design of frameworks could increase the bond strength of

resin composite to the fiber framework when chewing (Monaco et al 2003). The

fiber-reinforced materials have higher bending tendencies because of their lower

modulus compared to the more brittle ceramic material, but when the ceramic

material is supported by the zirconia framework the values can widely increase above

the threshold of the 500N. The ceramic fractures occur at the connectors between

pontic and abutment, where embrasure contours are critical. It should be kept in mind

that all of the results reported herein are from an in vitro study. Clinical conditions

may be more critical than those simulated in vitro and potentially may result in lower

forces creating failure. Despite of these potential limitations, however, the results of

this study support further investigation, and the clinical application of these materials

seems possible and promising.

References

Freilich MA, Karmaker AC, Burstone CJ, et al: Development and clinical applications of a light-polymerized fiber reinforcedcomposite. J Prosthet Dent 1998;80:311-318

Pauli C: Fracture strength of ceramometal and all-ceramic posterior fixed partial dentures. ZWR 1996;11:626-632

Rammelsberg P, Behr M, Pospiech P, et al: Adhesively fixed partial dentures: Esthetic and substance preserving alternatives for conventional fixed partial dentures. DZZ 1995;50: 224-227

Behr M, Rosentritt M, Lang R, et al: Flexural properties of fiber reinforced composite using a vacuum/pressure or a manual adaptation manufacturing process. J Dent 2000;28: 509-514.

Behr M, Rosentritt M, Latzel D, et al: Comparison of three types of fiber-reinforced composite molar crowns on their fracture resistance and marginal adaptation. J Dent 2001;29:187-196

Krejci I, Reich T, Lutz F, et al: In vitro test results of the evaluation of dental restoration systems. Correlation with in vivo results. Schweiz Monatsschr Zahnmed 1990;100:1445-1449.

Rosentritt M, Leibrock A, Lang R, et al: Regensburger masticator. Mater Test 1997;39:77-80

Delong R, Douglas WH: Development of an artificial oral environment for the testing of dental restoratives: Bi-axial force and movement control. J Dent Res 1983;62:32-36

55

Hidaka O, Iwasaki M, Saito M, et al: Influence of clenching intensity on bite force balance, occlusal contact area and average bite pressure. J Dent Res 1999;78:1336-1344.

Behr M, Rosentritt M, Leibrock A, et al: In vitro study of fracture strength and marginal adaptation of fiber-reinforced adhesive fixed partial inlay dentures. J Dent 1999;27:163-168.

Monaco C, Ferrari F, DDS, Miceli GP, Scotti R. Clinical Evaluation of Fiber-

Reinforced Composite Inlay Fixed Partial Dentures. Int J Prosthodont 2003;16:319-

325.

56

Chapter 5: laboratory process for high volume fiber framework

Missing single-tooth situations offer several reconstructive treatments modalities.

Adhesive bonded restorations made with fiber-reinforced composite (FRC) have

been proposed for the fabrication of metal-free inlay fixed partial dentures (IFPDs).

Nevertheless, for a small sized construction, such as a dental prosthesis, the design

and characteristics of the fiber framework are important and demand careful

attention. Fiber reinforcement should be optimal when the designing of the

prostheses and their components need to be realized to withstand masticatory

loading. This technical procedure describes the way to realize the cusp-supporting

framework with high volume of fibers and completely support the layering material.

57

The IFPD is indicated in all clinical cases were the replacement of a single tooth is

required but treatment with a dental implant cannot be realized. In this situation, an

inlay bridge offers a more conservative method of tooth replacement compared to a

crown retained conventional bridge because tooth preparations are limited to the

occluso and mesio/distal surfaces of abutment (Fig.1). However these preparations

are more clinical sensitive technique and required a careful laboratory procedure

(Monaco et al 2003). The cavity preparation must give the space to the fiber

framework that must completely support the pontic element. Freilich (Freilich at al

2002) hypothesized the increased rigidity and a broader base of support provided by

the FRC substructure was needed to support the composite veneer. Thus, they added

a substantial amount of FRC bulk to the pontic component of the substructure (low-

volume design), resulting in the creation of the “high-volume” substructure design

and examined their relationship with the clinical performance.

This technical procedure describes the realization of the fiber framework with an

anatomical shape and high volume of fibers that allows the extension of the pontic

element to support the layering material. The final design of the bridge’s framework

is similar to metal framework with extensions in the vestibular and buccal side to

completely hold up the veneer composite and avoid the cohesive fracture of the

composite.

Laboratory Technique

The glass fiber framework (Vectris; Ivoclar Vivadent) is made with pre-impregnated

‘pontic’ and ‘frame’ fibers with a monomer consisting of dimethacrylates and the

final sub-structure has an anatomical shape of the framework and a high volume of

the glass fibers.

The design of the fiberglass framework is first pre-modeled by a photo-curing resin

(Spectra Tray, Ivoclar Vivadent) to obtain the oval shape and its thickness is checked

in the molding model. The dental technician must empty the space for the veneering

material and check with carefulness the space of the connector surface. This area

must be minimum 9mm2 (3x3m), otherwise the flexibility of the framework could

increase and determine the worsening of the marginal adaptation. The photo-curing

resin framework is embedded in a transparent silicone impression paste to form a

mould. Then the resin model is removed and the glass fibers are inserted into the

58

silicone-mould. The pre-impregnated pontic fibers are condensed into the desired

shape by a vacuum- forming process and then cured by light in a VS1 unit (Ivoclar

Vivadent) for 10 min. According to the manufacturer’s recommendations, the pontic

fibers are treated with silane to increase the bonding to the following fiber layer. A

sheet of weave fibers (frame) is placed upon the pontic structure and the cycle in VS1

is repeated. After the moulding of the sheet fibers all excesses are removed with a

carbide bur and the framework is tried on the master model (Fig. 2). The veneering

material is built incrementally using a Quick light-curing unit. Finally, the IFPD is

placed into a Lumamat 100 unit (Ivoclar Vivadent) for the final application of light

and heat (104°C) to complete polymerization and maximize strength and other

physical characteristics (Fig. 3). After the polishing the IDPD is ready for the luting

procedures (Fig. 4).

Discussion

Many FRC researches on position and orientation have focused upon the effects of

the question of fiber reinforcement directionality like random, longitudinal

orientations or weaved. It is widely accepted that directional orientation of the fiber

long axis perpendicular to an applied force will result in strength reinforcement.

Forces that are parallel to the long axis of the fibers, however, produce matrix-

dominated failures and consequently yield little actual reinforcement. Design

strategies are on occasion employed to provide multi-directional reinforcement, to

minimize the highly anisotropic behavior of unidirectional fiber reinforcement ( Dyer

et al 2004). A higher volume fraction of fibers in the resin matrix improves the

mechanical properties. Moreover, the flexibility of the beam is in direct relation to

the amount, position and the type of the fibers that go to make up the framework. The

position of the FRC layer had an effect on the flexural strength of the framework.

The highest flexural strength was achieved when the FRC layer was located at the

tension side of the test specimens. The FRC structure benefits most when the tensile

stresses can be transferred to the reinforcing fibers (Vallittu 1998).

This laboratory technique allows the extension in the vestibular and buccal side to

support the veneering material like in metal or zirconia frameworks and hold it in all

loading directions. Vestibular and buccal FRC extensions in the pontic element can

59

increase the bonding area between the framework and resin composite as well as hold

the veneer material better during occlusal loading.

Conclusion

The design of the FRC framework can play an important role in supporting the

layering material in order to optimize the position and orientation of fibers that

should maximize the stress transferred from matrix to fibers and increase mechanical

properties of IFPDs.

Fig. 1 The presence of old restorations

on the abutments near the edentulous

space is the typical situation for the

IFPDs

Fig. 2 The anatomical fiber framework is

designed to completely support the

veneering composite and increase the

mechanical properties of the IFPD.

Fig. 3 . IFPD made with fiber reinforced

composite.

Fig.4 Follow-up after 2 years.

References

Dyer SR, Lassila LVJ, Jokinen M, Vallittu PK. Effect of fiber position and

orientation on fracture load of fiber-reinforced composite. Dent Mater 2004;20:

947–955.

Freilich MA, Meiers JC, Duncan JP, Eckrote KA, Goldberg AJ. Clinical

evaluation of fiber-reinforced fixed bridges. J Am Dent Assoc 2002;133:1524-

1534.

Monaco C, Ferrari F, DDS, Miceli GP, Scotti R. Clinical Evaluation of Fiber-

60

Reinforced Composite Inlay Fixed Partial Dentures. Int J Prosthodont 2003;16:319-

325.

Vallittu PK. Some aspects of the tensile strength of undirectional glass fibre-

polymethyl methacrylate composite used in dentures. J Oral Rehabil. 1998

Feb;25(2):100-5.

61

Chapter 6: Different structure of the framework

The traditional treatment for a single tooth replacement is a conventional fixed partial

denture (FPD). This technique requires a full-coverage preparation of the abutments.

Consequently, a large quantity of sound tooth is destroyed during the preparation

(Creugers et al 2000). This is particularly problematic in healthy and young teeth

with extensive pulp. In addition, the mean of a FPD is estimated at 8.3-10.3 years

(Creugers et al 2000, Näpänkangas et al 2002). Furthermore, a young patient would

require numerous replacements of this restoration over a lifetime. In order to limit

this destruction and thanks to evolution of adhesive dentistry, (Perdigão et al 1999)

resin bonded fixed partial dentures (RBFPDs) (Freilich et al 1998) and dental

implants (Leal et al 2001) constitute the current alternatives. These treatments have

several advantages over conventional bridges, especially in relation to conservation

of tooth structure and their reversibility (Hussey et al 1996).

Nevertheless, published data on the survival of RBFPDs with metal framework are

still controversial. Some study reports that resin-bonded bridges placed in the

posterior region were less retentive that FPDs placed in the anterior region (Creugers

et al 1997, Creugers et al 1992). De Kanter (De Kanter et al 1998) reported only 46%

and 62% of primary survival rates at 5 years of RBFPDs respectively for

conventional and modified preparation forms. A multi-centers study (Kerschbaum et

al 1996) reported a success rate of 66% that increased to 82% with additional

rebonding.

However, tooth-colored framework and a metal-free prosthetic reconstruction cannot

be built with this technique. Consequently, alternative materials such as high-strength

pressed ceramic (Sorensen et al 1999) and fiber-reinforced composites (FRC) have

achieved (Freilich et al 1998) a certain degree of popularity in restorative dentistry

today. Various types of fibers such as carbon fiber, aramid fiber and ultra-high

modulus polyethylene fiber have been tested (Ekstrand et al 1987, Gutterridge 1992,

Uzun et al 1999, Larson et al 1991, Hess et al 1987) as reinforcing materials, and it

has been shown that the fibers also increase the elastic modulus and flexural strength

of dental polymer. Recently, glass fibers have been used to reinforce denture base

polymers (Krause et al 1989). Glass fibers are favorable for denture base polymer

62

when used in light curing type resin because they have excellent transparency

compared to the other fibers (Vallittu 1999a). The advantages of FRC restorations are

that the laboratory procedure is simple because casting is not necessary, and the

bonding ability between resin and reinforced fiber is superior to that between resin

and casting metal, although surface treatment of fiber is necessary to obtain a strong

bonding (Vallittu 1999b). Reinforcement of dental resins with short or long fibers has

shown positive results and, generally, better mechanical properties when compared to

metal wires in polymers reinforcement (Vallittu 1996c). A glass FRC system made of

prefabricated glass fiber elements along with processing and curing equipment

(Targis-Vectris, Ivoclar-Vivadent) was introduced in 1996 (Krejci et al 1998). A

recent in vitro study reported high fracture strength of 700 N of fiber-reinforced

adhesive IFPDs. With one exception, the break always ran along the interface

between fiber-composite or in the Targis material itself (Behr et al 1999).

The framework in the FRC is supposed to be made in an oval shape in the pontic

element with parallel and weave fiberglasses. This laboratory technique allows the

extension in the vestibular and buccal side to support the layering material like in

metal framework and hold it in a continuous manner. In the present study, the

frameworks was made with different uses and shapes of the fiberglasses to

investigate the clinical behavior of IFPDs.

The purpose of this study was to collect survival data on posterior IFPDs, which were

placed under controlled clinical conditions and to identify the correlation between the

design of fiber frameworks and survival of IFPDs.

Materials and methods

Thirty patients were selected and received an indirect restoration. The selection of the

patients followed two criteria: their refusal to treatment with dental implants and

informed written consent. Selection of male and female subjects was restricted to

those aged 18-60 years and in good general and periodontal health. Patients with the

following factors were excluded from the clinical trial: patients receiving drugs that

modify pain perception, pregnancy or breast feeding, eating disorders, periodontal

surgery, well-known allergy to chemical compounds used in this study and active

periodontal disease.

63

From January 1998 to January 2002 forty-one IFPDs were inserted. Average service

time of Targis-Vectris was 27.2 months. Twenty-four restorations were placed in

mandibular teeth and 17 in maxillary teeth and evaluated using the USPHS criteria.

According to the manufacturer’s instructions the Vectris frameworks in the IFPDs

are made with pre-impregnated ‘pontic’ and ‘frame’ fibers (Ivoclar: Scientific

Document 1996). The FRC framework was made with only pontic fiber in nineteen

restorations (Group 1) and with pontic and frame pre-impregnated fiberglass in

twenty-two restorations (Group 2) according to the manufacturer’s instructions. In

the Group 1 the design of pontic element was made in a cylindrical shape whereas in

the Group 2 was modified in oval shape to hold the layering material in a continuous

manner (Fig 1). The final shape of Group 2 was similar to metal framework with

extensions in the vestibular and buccal side to support completely the veneer

composite.

Clinical Preparation Technique

At baseline all patients were tested for dentin sensitivity. In the rank order data, a

score of 0 was defined as no pain, 1-4 as mild sensitivity, which was provoked by the

dentists’ air blast, and 5-10 as strong sensitivity, which was spontaneously reported

by the patient during drinking or eating. The same measurement was performed again

at each recall. The status of the gingival tissues adjacent to the test sites was observed

at baseline and each recall.

Before cavity preparation a rubber dam was placed and all cavities were prepared

according to modified principles for adhesive inlays retainers to obtain adequately

strong dental restorations. No additional bonding FRC wings were made to obtain

more adhesional surface. The cavities were prepared with 80µm (No. 8113R, No.

8113NR, Intensiv SA) and finished with 25µm (No. 3113R, No. 3117, Intensiv SA)

diamond burs in a medium-speed handpiece with a water spray. All the dentin

surfaces of the cavity preparations were sealed with a three-component enamel-

dentin adhesive system (Syntac, Ivoclar-Vivadent). Undercuts and deep parts of the

cavities were covered with a resin composite (Tetric ceram, Ivoclar-Vivadent)

polymerized for 40 sec. All cavity margins were in enamel and extrasulcus. The

distance to marginal gingival was at least 1 mm. Complete mandibular and maxillary

64

arch impressions were taken with a polyvinyl siloxane (Imprint II, 3M ESPE). A

light-cured resin (Fermit-N, Ivoclar-Vivadent) was used as temporary restoration.

Laboratory Technique

The design of the fiberglass framework was first pre-modeled by a photo-curing resin

(Spectra Tray, Ivoclar) to obtain the oval shape and its thickness checked onto the

molding model. This model was embedded in a transparent silicone impression paste

to form a mould. Then this resin was removed and the fibers were applied into the

silicone-mould (Fig 2). The pre-impregnated ‘pontic’ fibers were condensed in a

deep-drawing, polymerization process. After a cycle of vacuum packing and heat in

VS1 unit (Ivoclar-Vivadent) for 9.30 min. according to the manufacturer’s

recommendations, the FRC was sandblasted with CoJet system (3M ESPE) with

small grain size of 30�m at 2,5 bar of pressure for 10 seconds or roughened with

8o�m diamond bur and treated with silane (Wetting agent, Ivoclar-Vivadent). A

sheet of wave fibers ‘frame’ was placed upon the ‘pontic’ structure and the cycle in

VS1 was repeated. The Targis material was built incrementally using Targis Quick.

Finally, the IFPD was placed into the Targis Power unit (Ivoclar-Vivadent) for the

final application of light and heat to complete polymerization and maximize strength

and other physical characteristics.

Adhesive Procedure

The IFPDs restorations were definitively inserted within 2 weeks after the impression

was made. The operating field was isolated with a rubber dam, provisional

restorations were removed and prepared teeth were cleaned with rubber cup and

pumice slurry (Fig 3). Restoration surfaces were treated with a 25�m diamond bur

(n.3274, Intensiv) in a medium-speed handpiece and a silane-coupling agent

(Monobond S, Ivoclar-Vivadent) was applied on the treated surfaces. All enamel and

dentin surfaces were treated with Syntac (Ivoclar-Vivadent) according to the

manufacturer’s instructions. The dual-cured composite resin cement Variolink II

(Vivadent), was applied to the inlays and the inside of the cavities with a disposable

brush. The luting cement was light activated for 60 sec. each from cervical, buccal,

lingual and occlusal surfaces. Occlusion and articulation were carefully checked after

65

the cementing step. The restorations were then finished with 15 µm diamond burs

(Composhape, Intensiv) and polished with a composite finishing and polishing kit

(Hawe Neos Dental) and Occlubrush (Hawe Neos Dental) in a slow-speed handpiece.

Evaluation

One investigator evaluated all the restorations directly after the final polishing (Fig 4)

and after 6, 12, 24 and 48 months. During the recalls, the patients came back to the

office to answer questions about post-operative sensitivity. Partial or total debonding

of IFPDs and framework or resin composite fractures were considered failures.

Patients with failure were examinated immediately to verify the assumption and the

precise time of the failure. The restorations were evaluated, using the USPHS

modified parameters, to check their stability and longevity in regard to the following

characteristics: color match, marginal discoloration, secondary caries, surface

texture, marginal adaptation, fracture and postoperative sensitivity. The scores were

Alfa (ideal), Bravo (clinically acceptable), Charlie (clinically unacceptable) scores.

Statistical Analysis

Statistical analysis was applied to compare the restorations at baseline and after last

recall, and to check for differences between Group 1 and 2. The Wilcoxon’s

matched-pairs signed ranks test and McNemar test measured the restorations’ success

at the appropriate time intervals and were used to rate all parameters. The Chi-square

test was used to examine for a significant difference in fracture rate between IFPDs

with conventional fiber framework structure and those with modified. The null

hypothesis was rejected at the 5% level. The survival estimation method of Kaplan-

Maier was used with statistical software (Primer 4.02, Stanton A. Glantz, McGraw-

Hill).

Results

No partial/total debonding of the IFPDs or breakage of fiber frameworks were shown

during the observation period. Three partial adhesional/cohesional veneering

composite fractures (Fig 5) occurred in the occlusal-buccal side of the pontic element

in Group 1 after 3, 4 and 8 months. Intraoral repairs were made in all cases of

66

breakage and two IFPDs were completely replaced after a few weeks because the

composite repairs had debonded. After each failure, the fiber framework was visible

at low magnification (Zeiss 3.6x35mm). One cohesive fracture of Targis (Fig 6)

occurred in the abutment in Group 2 after 46 months. The FPD was repaired after

two weeks of intraoral exposion and it is still in place. No statistical differences

regarding fractures were showed between baseline and last recall and between Group

1 and 2 (P>0.05). The percentage of Bravo color match for the IFPDs increased from

7% at baseline to 29% at last recall with statistically significant difference (P<0.048).

The small changes noted during the follow-up period in clinical results for the

abutments were not statistically significant (Table 1). Four patients reported

postoperative hypersensitivity. These patients reported temperature sensitivity and

chewing pain for periods lasting from 1 week to 2 months. These complaints

disappeared completely within 6 months. The Kaplan-Maier survival estimation was

86% for Group 1 and Group 2 (Fig 7).

Table 1 Clinical Results Based on USPHS Criteria

Abutments (n=82)

Baseline

Alfa

Last Recall

Alfa

Baseline

Bravo

Last Recall

Bravo

% Alfa at

Last

Recall

Marginal discoloration 82 79 0 3 96

Secondary caries 82 81 0 1 99

Marginal adaptation 82 80 0 2 98

Postoperative sensibility 78 82 4 0 100

IFPDs (n=41)

Fracture 41 37 0 4 90

Surface texture 41 36 0 5 88

Color match 38 29 3 12 71

Discussion

Inlay fixed partial dentures with fiber framework offer a new option to restore a

missing tooth with a less invasive method and with better esthetic treatment than with

67

metal FPDs. Data on the clinical performance of FRC is needed to estimate the

survival rates as well as to gain knowledge on the appropriate use of fibers in

combination with resin composites. There are few clinical studies on the behavior of

FPDs and most of them are short-term evaluations (Göhring et al 1999, Vallittu et al

2000, Lutz et al 2000). One of the advantages of IFPDs is their intraoral repairing

(Swift et al 1992) but post light curing and heat treatment determine a high degree of

conversion. This treatment improves mechanical properties but it makes the repair of

IFPDs more difficult.

In our study, all adhesional/cohesional veneering composite fractures occurred in the

pontic element in Group 1. Similar failures were reported in an earlier experimental

study with fiber-reinforced composite using a laboratory multistep (Altieri et al

1994). Two of the three intraoral repairs failed after 2 weeks. There are a few

plausible explanations for the failures.

First, the design of the FRC framework can play important role in supporting the

layering material in a continuous manner. Vestibular and buccal FRC extensions in

the pontic element can increase the bonding area between the framework and resin

composite as well as hold the veneer material better during occlusal loading. Failure

was invariably seen in the same position in all bridges with conventional framework

design. This observation suggested that modified design of frameworks could

increase the bond strength of resin composite to the fiber framework when chewing.

Continuous unidirectional fibers gave the highest strength and stiffness only in the

direction of the fiber. Woven fibers are able to reinforce the denture base polymers in

two directions (DeBoer et al 1984). However, definitive recommendations for

prostheses manufacturing currently cannot be made. Further in vitro and clinical

studies must be conducted to confirm this hypothesis.

Second, there might have been inadequate bonding between fiber framework and

composite especially after repair. Rosentritt (Rosentritt et al 1998) demonstrated that

median facing fracture strength of Targis/Vectris bridges was 1450 N after a

simulated clinical service of 5 years but the fracture strengths of the repair IFPDs

were significantly lower compared with the control group after an additional

simulated 2-year interval. However, the median fracture force was still 1000 N

higher than chewing force. Dynamic tests would have different value in comparison

to the static tests. Furthermore, they showed that aluminum oxide pretreatment or

68

abrading with diamond burs and silanization provided in vitro sufficient fracture

strength. Nevertheless, an aqueous environment can induce corrosive effects on the

surface of glass fibers resulting from water absorption through the polymer matrix

(Lassila et al 2002). Other substances, like alcohol, can break the polymer chains

with a softening effect on the composite that can lead to a reduction of the

mechanical strength (Ehrenstein et al 1990). The bonding values between a repair

and an aged framework are about 20% to 70% lower in comparison to the bonding

between a new framework and veneer (Swift et al 1994). Therefore, the design of the

IFPD after repair should help to preserve points of high stress, especially in occlusal

areas.

Third, it is possible that cohesional fractures inside resin composite may be caused

by air entrapment during manufacturing to ensure esthetics. Entrapment of air locally

inhibits polymerization and weakens the structure of materials (Vallittu et al 1997).

Görhing (Görhing et al 1999) described a modified manufacturing technique to

eliminate occlusal multilayering and to reduce the risk of veneering composite

fractures. The matrix chemistry of the repair composite might not influence the

repair strength (Gregory et al 1990), but the composite should provide low viscosity

to achieve a good wetting of the repair point (Rosentritt et al 1997). The moderate

decrease in fracture resistance obtained after veneering repair in vitro of

Targis/Vectris bridges indicates that repeated repair is possible (Rosentritt et al

2001), although it is highly risky to repair an extremely damaged FPD that has lost

more than 50% of the veneer substance.

The percentage of bravo in color match was 29% at last recall and the deterioration

was significant compared to initial status. Some inlay FPDs lost their surface shine

with the color that was slightly too dark/opaque with superficial stains and the

repolishing was only partially successful in reversing the discoloration. Color

changes or resin material can occur through several mechanisms: formation of

degradation products changes in surface structure due to the wear and by extrinsic

staining (Powers et al 1980). Douglas (Douglas 2000) showed that after accelerated

aging the new generation of indirect resins demonstrated color changes at or below a

quantitative level that would be considered acceptable. These changes were in the

magnitude of 0.62 to 3.4�E units. However, Stober (Stober et al 2001) described

changes in color of the same composite after 4 and 8 week storage in various

69

substances generally thought to cause intra-oral discoloration. In particular, red wine

and turmeric solution caused severe discoloration with total differences of ∆E >10 in

all tested composites. It is apparent from our results that it remains essential to

improve the color stability of Targis material used for esthetic facing. In a short-term

observation period, the FRC showed good clinical service but these results need more

observation in order to create a sound basis for the final assessment of this method.

In our study the veneering composite had less fractures with modified FRC

frameworks than with the conventional design. Based on our observations the repair

of FRC fractured veneer is advisable only for small damages.

References

Altieri JV, Burtone CJ, Goldberg AJ, Patel AP. Longitudinal clinical evaluation of fiber-reinforced composite FPDs: a pilot study: J Prosthet Dent 1994;71:16-22.

Behr M, Rosentritt M, Leibrock A, Schneider-Feyrer S, Handel G. In vitro-study

of fracture strength and marginal adaptation of fibre-reinforced adhesive fixed partial inlay dentures. J Dent 1999;27:163-168.

Creugers NH, De Kanter RJ, Van’t Hof MA. Long-term survival data from a

clinical trial on resin bonded bridges. J Dent 1997;25:239-242. Creugers NH, Käyser AF, Van’t Hof MA. A seven-and-half–year survival study

of resin-bonded bridges. J Dent Res 1992;71:1822-1825. Creugers NHJ, De Kanter RJ. Patients’ satisfaction in two long-term clinical

studies on resin-bonded bridges. J Oral Rehabil 2000;27:602-607. Creugers NHJ, Käyser AF, Van’t Hof MA. A meta-analysis of durability data on

conventional fixed bridges: Community Dent and Oral Epidemiol 1994; 22:448-452. De Kanter RJ, Creugers NH,Verzijden CW, Van’t Hof MA. A five-year multi-

practice clinical study on posterior resin-bonded bridges. J Dent Res 1998;77:609-614.

DeBoer J, Vermilyea SG, Brady RE. The effect of carbon fiber orientation on

the fatigue resistance and bending properties of two denture resins. J Prosthet Dent 1984;51:119-121.

70

Douglas DR. Color stability of new-generation indirect resins for prosthodontics application. J Prosthet Dent 2000;83:166-170.

Ehrenstein GW, Schmiemann A, Bledzki A, Spaude R. Corrosion phenomena in

glass-fiber-reinforced thermosetting resins. In: Cheremisinoff NP (ed). Handbook of ceramics and composites, vol 1. New York: Marcel Dekker, 1990:231-268.

Ekstrand K, Ruyter IE. Carbon/graphite fiber reinforced poly(methyl

methacrylate): properties under dry and wet conditions. J Biomed Mater Res 1987;21:1065-1080.

Freilich MA, Duncan JP, Meiers JC, Goldberg AJ. Preimpregnated, fiber-

reinforced prostheses. Part I: basic rationale and complete coverage and intracoronal fixed partial denture designes. Quintessence Int 1998;29:689-696.

Freilich MA, Duncan JP, Meiers JC. Preimpregnated, fiber-reinforced

prostheses. Part I. Basic rationale and complete-coverage and intracoronal fixed partial denture designs. Quintessence Int 1998:29:689-696.

Göhring TN, Krejci I, Lutz F. Adhäsive Inlaybrücken aus glasfaserverstärktem

Komposit. Step by step-Beschreibung der klinischen Anwendung. Schweiz Monatsschr Zahnmed 1999;109:369-379.

Göhring TN, Mörmann WH, Lutz F. Clinical and scanning electron microscopic

evaluation of fiber-reinforced inlay fixed partial dentures: Preliminary results after one year. J Prosthet Dent 1999;82:662-668.

Gregory DC, Pounder B, Bakus E. Bond strengths of chemically dissimilar

repaired composite resins. J Prosthet Dent 1990;64:664-668. Gutterridge DL. Reinforcement of poly(methyl methacrylate) with ultra-high-

modulus polyethylene fiber. J Dent 1992;20:50-54. Hess D, Belser U. Provisional extension bridges reinforced with Kevlar.

Schweiz Monatsschr Zahnmed 1987;97:457-463. Hussey DL, Linden GJ. The clinical performance of cantilevered resin-bonded

bridgework. J Dent 1996;24:251-256. Ivoclar: Scientific Document. Ivoclar-Vivadent, Schaan, FL, 1996. Kerschbaum T, Haastert B, Marinello CP. Risk of debonding in three-unit resin-

bonded fixed partial dentures. J Prosthet Dent 1996;75:248-253. Krause WR, Park SH, Straup RA. Mechanical properties of BIS-GMA resin

short glass fiber composites. J Biomed Mat Res 1989;23:1195-1211. Krejci I,Boretti R, Giezendanner P, Lutz F. Adhesive crowns and fixed partial

dentures fabricated of ceromer/FRC: clinical and laboratory procedures Pract

71

Periodont Aesthet Dent 1998;10:487-498. Larson WR, Dixon DL, Aquilino SA, Clancy JM. The effect of carbon graphite

fiber reinforcement on the strength of provisional crown and fixed partial denture resins. J Prosthet Dent 1991;66:816-20.

Lassila LVJ, Nohrström T, Vallittu K. The influence of short term water storage

on the flexural properties of unidirectional glass fiber-reinforced composite. Biomaterials 2002;23:2221-2229.

Leal FR, Cobb DS, Denehy GE, Margeas RC. A conservative aesthetic solution

for a single anterior edentulous space: case report and one-year follow-up. Pract Proced Aesthet Dent 2001;13:635-641.

Lutz F, Göhring TN. Fiber-reinforced inlay fixed partial dentures: maximum

preservation of dental hard tissue. J Esthet Dent 2000;12:164-171. Näpänkangas R, Salonen-Kemppi MAM, Raustia AM. Longevity of fixed metal

ceramic bridge prostheses: a clinical follow-up study. J Oral Rehabil 2002;29:140-145.

Perdigão J, Lopes M. Dentin bonding-state of art 1999. Compend Contin Educ

Dent 1999;20:1151-1162. Powers JM, Fan PL, Raptis CN. Color stability of a new composite restorative

material under accelerated aging. J Dent Res 1980;59:2071.2074. Rosentritt M, Behr M, Kolbeck C, Handel G. In vitro repair of three-unit fiber-

reinforced composite FPDs. Int J Prosthodont 2001;14:344-349. Rosentritt M, Behr M, Leibrock A, Handel G, Friedl KH. Intraoral repair of

fiber-reinforced composite fixed partial dentures. J Prosthet Dent 1998;79:393-398. Rosentritt M, Leibrock A, Lang R, Behr M, Scharnagl P, Handel G.

Regensburger masticator. Mater Test 1997;3:77-80. Sorensen JA, Cruz M, Mito WT, Raffeiner O, Meredith HR. Foser HP. A

clinical investigation on three-unit fixed partial dentures fabricated with lithium disilicate glass-ceramic. Pract Periodontics Aesthet Dent 1999;11:95-106.

Stober T, Gilde H, Lenz P. Color stability of highly filled composite resin

materials for facings. Dent Mater 2001;17:87-94. Swift EJ, Le Valley BD, Boyer DB. Evaluation of new methods for composite

repair. Dent Mater 1992 ;8:362-5. Swift, EJ, Bryan CC, Boyer DB. Effect of silane coupling agent on composite

repair strengths. Am J Dent 1994;7:200-202. Uzun G, Hersek N, Tincer T. Effect of five woven fiber reinforcements on the

72

impact and transverse strength of a denture base resin. J Prosthet Dent 1999;81:616-620.

Vallittu PK, Sevelius C. Resin-bonded, glass fiber-reinforced composite fixed

partial dentures: a clinical study. J Prosthet Dent 2000; 84:413-418. Vallittu PK. A review of fiber-reinforced denture base resin. J Prosthodont

1996;5:270-276. Vallittu PK. Flexural properties of acrylic resin polymers reinforced with

unidirectional and woven glass fibers. J Prosthet Dent 1999;25:100-105. Vallittu PK. Oxygen inhibition of autopolymerization of

polymethylmethacrylate-glass fiber composite. J Mater Sci Mater Med 1997;8:489.492.

Vallittu PK. Prosthodontic treatment with a glass fiber-reinforced resin-bonded

fixed partial denture: a clinical report. J Prosthet Dent 1999;82:132-135.

73

Chapter 7: adhesive procedures

7.1 Inlay Bridge With a New Microfilled Composite: A Clinical Report

74

The traditional treatment for a single tooth replacement is a conventional fixed

partial denture (FPD). This technique requires a full-coverage preparation of the

abutments. Consequently, a large quantity of sound tooth is destroyed during the

preparation (Creugers et al 2000). This is particularly problematic in healthy and

young teeth with extensive pulp. In addition, the mean of a FPD is estimated at 8.3-

10.3 years (Creugers et al 1994, Näpänkangas et al 2002). Thanks to evolution of

adhesive dentistry (Perdigão 1999) this destruction can be limited. Resin bonded

fixed partial dentures (RBFPDs) (Freilich et al 1998) and dental implants (Leal et al

2001) constitute the current alternatives. These treatments have several advantages

over conventional bridges, especially in relation to conservation of tooth structure

and their reversibility (Hussey et al 1996).

Fiber-reinforced composites (FRC) have achieved (Freilich 1998) a certain degree of

popularity in restorative dentistry today (Vallittu 1999a). Glass fibers are favorable

for denture base polymer when used in light curing type resin because they have

excellent transparency compared to the other fibers. The advantages of FRC

restorations are that the laboratory procedure is simple because casting is not

necessary, and the bonding ability between resin and reinforced fiber is superior to

that between resin and casting metal, although surface treatment of fiber is necessary

to obtain a strong bonding (Vallittu 1999b).

A novel type of composite material (SR Adoro; Ivoclar Vivadent) has been recently

introduced for dental use. One advantage of this new material is the ability to form

good bonding with the polymer matrix of the glass fiber system (Vectris; Ivoclar

Vivadent) and the luting composite.

This clinical report describes the treatment of a patient with a missing lower first

molar with a new microfilled composite supported with a glass fiber framework.

Inlay Bridge indication

The inlay bridge is indicated in all clinical cases were the replacement of a single

tooth is required but treatment with a dental implant cannot be realized. In the

following situations treatment with implants may not be possible: patient rejects

surgical treatment, costs of implant therapy are too high, not enough bone remaining

75

for the placement of an implant, poor prognosis of implant therapy due to patient

related situations, e.g., smoking and insufficient oral hygiene habits. In this situation,

an inlay bridge offers a more conservative method of tooth replacement compared to

a crown retained conventional bridge because tooth preparations are limited to the

occluso and mesio/distal surfaces of abutment. However these preparations are more

technique sensitive and required a careful adhesive luting procedure. Short spans, not

exceeding 10-12 mm (the space of a molar) are recommended and the preparation

must give the space to the fiber framework to support completely the pontic element.

Clinical Report

A 35-year-old woman presented herself in the Department of Prosthetic Dentistry

of University of Bologna for the replacement of the first lower molar. The patient

was previously treated by her dentist with implant technique without success and was

not willing to have another surgical treatment because of the additional expense.

Intraoral examination revealed the presence of OD composite restoration on the

second premolar and an amalgam filling with a medium decay on the second molar.

The patient opted for a treatment with a conservative glass fiber composite RBFPD.

Clinical Preparation Technique

Before cavity preparation a rubber dam was placed and the cavities were prepared

according to modified principles for adhesive inlays retainers to obtain adequately

strong dental restorations.

No additional bonding FRC wings were made to obtain more bonding surface. The

cavities were prepared with 80 µm (No. 8113R, No. 8113NR, Intensiv SA) and

finished with 25 µm (No. 3113R, No. 3117, Intensiv SA) diamond burs in a medium-

speed handpiece with a water spray and ultrasonic points (SonicSys; Kavo). All the

dentin surfaces of the cavity preparations were sealed with an adhesive system

(Excite DSC, Ivoclar Vivadent). Undercuts and deep parts of the cavities were

covered with a resin composite (Tetric Flow, Ivoclar Vivadent) polymerized for 40

sec. All cavity margins were in enamel and extrasulcus. The distance to marginal

76

gingival was at least 1 mm. Complete mandibular and maxillary arch impressions

were taken with a polyvinyl siloxane (Imprint II, 3M ESPE). A light-cured resin

(Fermit-N, Ivoclar Vivadent) supported by glass fibers of Pontic (Ivoclar Vivadent)

was used as temporary restoration.

Adhesive Procedure

The IFPDs restorations were definitively inserted within 2 weeks after the

impression was made. The operating field was isolated with a rubber dam,

provisional restorations were removed, the teeth were cleaned Tubulicid and rubber

cup and pumice slurry and prepared inside restoration surfaces were sandblasted with

CoJet system (3M ESPE) with small grain size of 30 µm at 2 bar of pressure for 10

seconds. The inlay bridge was tried to check the fit of the restoration. The inner

surfaces of the inlay bridge were sandblasted with CoJet system (3M ESPE) with

small grain size of 30 µm at 2 bar of pressure for 10 seconds and a silane-coupling

agent (Monobond S, Ivoclar Vivadent). All enamel and dentin surfaces were treated

with Excite DSC (Ivoclar Vivadent) according to the manufacturer’s instructions.

The dual-cured luting composite Variolink II (Ivoclar Vivadent), was applied to the

inlays and the inside of the cavities with a disposable brush. The restoration was

inserted in the cavities with a stick and all excess of luting composite was removed.

A layer of a glycerin (Liquid Strip; Ivoclar Vivadent) was applied on the margin

before the polymerization (Fig. 119). The luting composite was light activated for 60

sec. each from cervical, buccal, lingual and occlusal surfaces. The restorations were

then finished with 15 µm diamond burs (Composhape, Intensiv) and polished with a

composite finishing and polishing kit (Hawe Neos Dental) and Astrobrush (Ivoclar

Vivadent) in a slow-speed handpiece. Occlusion and articulation were carefully

checked after the luting step.

Discussion

The replacement of missing tooth with a RBFPD is a very conservative

prosthodontic treatment especially when the patients refuse the treatment with

dental implants. This article describes the replacement of a molar with a

conservative inlay bridge resulting in the good success in the short-term follow up.

77

In a previous study on inlay bridge11 the FRC showed good clinical service but

these results need more observation in order to create a sound basis for the final

assessment of this method. In this study the veneering composite had less fractures

with anatomically shaped FRC frameworks which support the occlusal surface of

the pontic teeth than with the conventional design. The two important points to

achieve the success with this conservative technique are the cavity preparation and

the design of the fiber framework. The cavity must give the space to the fiber

framework to support completely the pontic element. With only very small

preparations such as slots debonding has been a problem. For the retention of Inlay

bridges Ivoclar Vivadent recommends at least two surface cavities of the following

extension: Depth of occlusal cavity ≥2.5 mm; approximo-central extension ≥4mm;

axial depth ≥1.2 mm; proximal step tooth and defect oriented, ideally 2 mm;

bucco-oral width ≥3.5 mm. Nevertheless, if the extension of the defect permits, we

try to realize a larger axial depth and adopted the following preparation guidelines.

Occlusal slot must be 2.0-2.5 mm, occlusal step 2.0 mm, proximal step 2.0 mm,

axial depth 2.0 mm and buccal-vestibular width 3.5 or greater.

The potential advantages of this technique are self-evident. First, the procedure can

be completed in two appointments with low-cost and preservation of sound tooth

tissue. Second, the periodontal apparatus of the abutment teeth is not compromised

in any way. Third, because this approach is relatively little invasive. It permits the

patient to opt for other, more traditional tooth replacement methods in the future.

Fourth, repairs can be carried out directly, without the need for any complicated

techniques or materials.

References

Creugers NHJ, De Kanter RJ. Patients’ satisfaction in two long-term clinical

studies on resin-bonded bridges. J Oral Rehabil 2000;27:602-607.

Creugers NHJ, Käyser AF, Van’t Hof MA. A meta-analysis of durability data on

conventional fixed bridges: Community Dent and Oral Epidemiol 1994; 22:448-

452.

78

Näpänkangas R, Salonen-Kemppi MAM, Raustia AM. Longevity

of fixed metal ceramic bridge prostheses: a clinical follow-up study.

J Oral Rehabil 2002;29:140-145. Perdigão J, Lopes M. Dentin bonding-state of art 1999. Compend Contin Educ

Dent 1999;20:1151-1162.

Freilich MA, Duncan JP, Meiers JC. Preimpregnated, fiber-reinforced prostheses.

Part I. Basic rationale and complete-coverage and intracoronal fixed partial

denture designs. Quintessence Int 1998:29:689-696.

Leal FR, Cobb DS, Denehy GE, Margeas RC. A conservative aesthetic solution

for a single anterior edentulous space: case report and one-year follow-up. Pract

Proced Aesthet Dent 2001;13:635-641.

Hussey DL, Linden GJ. The clinical performance of cantilevered resin-bonded

bridgework. J Dent 1996;24:251-256.

Freilich MA, Duncan JP, Meiers JC, Goldberg AJ. Preimpregnated, fiber-

reinforced prostheses. Part I: basic rationale and complete coverage and

intracoronal fixed partial denture designes. Quintessence Int 1998;29:689-696.

Vallittu PK. Flexural properties of acrylic resin polymers reinforced with

unidirectional and woven glass fibers. J Prosthet Dent 1999;25:100-105.

Vallittu PK. Prosthodontic treatment with a glass fiber-reinforced resin-bonded

fixed partial denture: a clinical report. J Prosthet Dent 1999;82:132-135.

79

Chapter 8: Clinical trial

8.2 Randomized controlled trial of Fiber-Reinforced Composite Inlay Fixed Partial

Dentures: two-year results.

80

The metal-ceramic fixed partial denture (FPD) is a standard restoration for a

single tooth replacement in dental practice. This porcelain-fused-to-metal (PFM) has

demonstrated an excellent record of service for many years, but continues to exhibit

several drawbacks (Freilich et al 1998). A significant amount of sound tooth structure

must be removed when fabricating conventional FPDs. Other potential disadvantages

of the conventional FPDs are discoloration of the gingival and visible metal margins

or “shine through” effects of the metal frameworks (Kolbeck et al 2002). Moreover

base metal alloys may exhibit corrosion and/or may elicit an allergic reaction from a

segment of the patient population (Council on Dental Materials, Instruments and

Equipment 1995).

Implant-supported restorations provide high-quality alternatives to conventional

FPDs. However, in some cases the treatment with implants may not be possible: the

patient rejects surgical treatment, the costs of implant therapy are too high, not

enough bone remaining for the placement of an implant (Sewon et al 2000), poor

prognosis of implant therapy due to patient related situations, e.g., smoking (Gruica

et al 2004) and insufficient oral hygiene habits (Quirynen et al 2002).

In these situations, an inlay fixed partial denture (IFPD) offers a more

conservative method of tooth replacement compared to a crown retained

conventional bridge because tooth preparations are limited to the occluso and

mesio/distal surfaces of abutment (Meiers et al 2001). However these preparations

are more technique sensitive and require a careful adhesive luting procedure (Vallittu

2004). Since the bonding procedures strengthen the cusps and provide additional

support for dentition, minimally invasive preparation is feasible (Morin et al 1984).

The alternatives to conventional PFM prostheses include all-ceramic, all-

particulate composite, and fiber-reinforced composite (FRC) systems. The all-

ceramic and all-particulate composite systems have been described for FPDs (Kern et

al 1991) and IFPDs (Beuchat et al 1999) but, in general, exhibit low resilience and

toughness and are subject to fracture (Fischer et al 2003, Zumbuhl et al 2000).

FRC is composed of two types of composite materials: fiber composites to build

the substructure and a hybrid or microfilled composite to veneer the external

surfaces. Initial in vitro investigations of FRC restorations demonstrated promising

results. After simulation of oral stresses, the fracture resistance and marginal

adaptation of adhesively fixed molar crowns, IFPDs and three-unit complete

81

coverage FPDs were better than for all-ceramic restorations (Loose et al 1998). The

fracture resistance of IFPDs showed a mean of about 700N, a value that led to the

expectation that these restorations would be successful under clinical conditions

(Behr et al 1999). Marginal adaptation of adhesively cemented FRC restorations was

shown to remain statistically unchanged after simulation of 5 years of oral stress. It

can be concluded from the in vitro wear investigations that the veneering composite

will have a wear rate comparable to enamel during a period of 5 years, and that it

can bear the load in occlusal contact areas (Kern et al 1999, Knobloch et al 1999).

After 2 years of observation, an investigation on forty IFPDs reported 89.6% of

continuous margin at the tooth-luting composite interface but four inlay bridges

failed for debonding or delamination of veneering material (Göhring et al 2002).

Similar results on the survival rate were obtained in two different studies after a

three-year observation period using two different FRC systems. Behr (Behr et al

2003) achieved 72% of survival rate using the Targis/Vectris (Ivoclar Vivadent Inc.,

Schaan, Liechtenstein) system as Freilich (Freilich et al 2002) that reported 75% of

survival rate for FiberKor/Sculpture (Pentron Laboratory Technologies LLC,

Wallingford, Conn). Fiber-reinforced composite used in IFPDs is a promising

material group but little clinical information is available compared to the traditional

prosthodontic materials.

The purpose of this randomized controlled trial was to collect survival data on

posterior IFPDs, using a new microfilled composite in combination with a fiber

framework system (SR Adoro/Vectris, Ivoclar Vivadent) placed under controlled

clinical conditions and to identify the clinical behaviour of two different bonding

systems used to lute the inlay bridges (Excite DSC versus Syntac, Ivoclar Vivadent).

The null hypothesis tested in this study was: there is no difference in postoperative

sensitivity and clinical behaviour between the two-step dual cured adhesive versus

the three-step adhesive bonding system in IFPDs.

MATERIALS AND METHODS

Thirty-nine patients were screened and received an indirect restoration. The selection

of the patients followed two criteria: their refusal to treatment with dental implants

and informed written consent. The study was approved by the Ethics Committee of

82

the University of Bologna. Selection of male and female subjects was restricted to

those aged 18-60 years and in good general and periodontal health. Patients with the

following factors were excluded from the clinical trial: patients receiving drugs that

modify pain perception, pregnancy or breast feeding, eating disorders, periodontal

surgery, well-known allergy to chemical compounds used in this study. The inclusion

criteria were: patients with a missing tooth, absence of any active perio or pulpal

disease, proximal margin located above the cementum-enamel junction, placement of

the rubber dam possible and the greatest distance between the abutments was smaller

or equal to 12 millimeters. The patients were divided in two groups. In the nineteen

patients of group A the Excite DSC adhesive system was used, in those of group B

the cementation was carried out with Syntac. The randomization of the patients was

performed with the toss of a coin.

From June 2002 to July 2004 thirty-nine IFPDs were inserted. Twenty-two

restorations were placed in mandibular teeth and 17 in maxillary teeth and evaluated

using the USPHS criteria. Three bridges had been luted to replace the first premolar,

fifteen for the second premolar and twenty-one to replace the first molar (Fig. 1-6).

Twenty-four restorations were between 2- and 3-year old, eleven bridges were

included between one and two years and only four restorations were less 1-year old

but had more than six months. Average service time of the IFPDs was 23.4 months.

Clinical Preparation Technique

At baseline all patients were tested for dentin sensitivity. In the rank order data, a

score of zero was defined as no pain (alpha), 1-4 as mild sensitivity (bravo), which

was provoked by the dentists’ air blast, and 5-10 as strong sensitivity (Charlie),

which was spontaneously reported by the patient during drinking or eating. The same

measurement was performed again at each recall. The status of the gingival tissues

adjacent to the test sites was observed at baseline and each recall.

Before cavity preparation a rubber dam was placed and all cavities were prepared

according to modified principles for adhesive inlays retainers to obtain adequately

strong dental restorations. No additional bonding FRC wings were made to obtain

more adhesional surface. The cavities were prepared with 80µm (No. 8113R, No.

8113NR, Intensiv SA) and finished with 25µm (No. 3113R, No. 3117, Intensiv SA)

83

diamond burs in a medium-speed handpiece with a water spray. The design of the

cavity preparations followed the philosophy of maximal preservation of sound tooth

(Göhring et al 1999). Pre-existing restorations were removed and their cavities were

used as abutments after appropriate preparation. In situations of primary carious

lesions, defect-oriented tooth cavities were prepared within the concept of minimal,

but adequately sized, inlay abutments. The cavity had to give the space to the fiber

framework to completely support the pontic element. The form of the cavity was at

least two surfaces with the following extension: depth of occlusal cavity ≥2.5 mm;

occluso-cervical height ≥4mm; axial depth ≥1.2 mm; bucco-oral width ≥3.5 mm. The

measurements of the cavities were verified with a periodontal probe after the

finishing of the cavity margins. Nevertheless, if the extension of the defect permits,

we try to realize a larger axial depth and adopted the following preparation

guidelines. A taper of ≈4° (or larger if given by the pre-existing restoration) was

chosen to simplify insertion. According to the pre-existing restorations, most

proximal boxes were larger. The distribution of the cavity preparations are described

in table 1.

All the dentin surfaces of the cavity

preparations were etched for 15” with

35% phosphoric acid and sealed with two

bonding systems Excite DSC or Syntac

(Ivoclar Vivadent). The same adhesive

system used for the dentin hybridization

was employed for the luting procedure.

Undercuts and deep parts of the cavities

were covered in increments with a highly

viscous composite (Tetric Ceram, Ivoclar

Vivadent) polymerized for 40 sec. The

excesses of the bonding on the enamel surfaces were removed with 25µm (No.

3113R, No. 3117, Intensiv SA) diamond burs in a medium-speed handpiece with a

water spray. All cavity margins were in enamel and extrasulcus. The distance to

marginal gingival was at least 1 mm. Complete mandibular and maxillary arch

Table 1 Different distribution

of the preparations

Slot 9

Mesio-occlusal 21

Disto-occlusal 23

Mesio-disto-occlusal 14

Onlay 5

Overlay 4

Crown 2

Total abutments 78

84

impressions were taken with a polyether material (Permadyne, 3M ESPE). A light-

cured resin (Fermit-N, Ivoclar Vivadent) was used as temporary restoration.

Laboratory Technique

According to the manufacturer’s instructions (Transil technique), the Vectris

frameworks in the IFPDs were made with pre-impregnated ‘pontic’ and ‘frame’

fibers and all restorations had an anatomical shape of the framework like an oval

shape to hold the layering material in a continuous manner and reach a high volume

of the substructure. The final design of the bridge’s framework was similar to metal

framework with extensions in the vestibular and buccal side to completely support

the veneer composite.

The design of the fiberglass framework was first pre-modeled by a photo-curing resin

(Spectra Tray, Ivoclar Vivadent) to obtain the oval shape and its thickness checked in

the molding model. This model was embedded in a transparent silicone impression

paste to form a mould. Then this resin was removed and the fibers were applied into

the silicone-mould (Transil, Ivoclar Vivadent). The pre-impregnated ‘pontic’ fibers

were condensed into the desired shape by a vacuum- forming process and then cured

by light in a VS1 unit (Ivoclar Vivadent) for 10 min. According to the manufacturer’s

recommendations, the FRC was treated with silane (Wetting agent, Ivoclar-

Vivadent). A sheet of wave fibers ‘frame’ was placed upon the ‘pontic’ structure and

the cycle in VS1 was repeated. The SR Adoro material was built incrementally using

a Quick light-curing unit. Finally, the IFPD was placed into a Lumamat 100 unit

(Ivoclar Vivadent) for the final application of light and heat (104°C) to complete

polymerization and maximize strength and other physical characteristics.

Adhesive Procedure

The IFPDs restorations were inserted within 2 weeks after the impression was

made. The operating field was isolated with a rubber dam, provisional restorations

were removed with a sharp probe and the prepared teeth were cleaned with nylon

bristle brushes and water spray. The inner surfaces of the inlay retainers were

sandblasted with CoJet system (3M ESPE) with small grain size of 30 µm at 2 bar of

pressure for 10 seconds. These treated inner surfaces were then silanated by using

85

Monobond-S (Ivoclar Vivadent). Just before the final cementations, the surfaces

were the bonding agent (Heliobond or Excite DSC; Ivoclar Vivadent) was brushed

on the surfaces and air thinned. To prevent early polymerization of this layer

especially for the Heliobond, the bridge was shielded against light until insertion. All

enamel finish lines were conditioned with 35% phosphoric acid gel (Ivoclar

Vivadent) for 30 seconds whereas the dentin surfaces were etched with the same acid

for 15 seconds. The Group A was treated with Excite DSC and Group B with Syntac

according to the manufacturer’s instructions. Both the adhesive bonding systems

weren’t polymerized before the placement of the luting composite. The bonding

agent was blown to a thin layer and the dual-cured composite resin cement Variolink

II (Ivoclar Vivadent) was used to lute the restorations. The luting composite was light

activated for 60 sec. (Optilux 500, Kerr) each from cervical, buccal, lingual, and

occlusal surfaces. Occlusion and articulation were carefully checked after the

cementing step. The restorations were then finished with 15�m diamond burs

(Composhape, Intensiv) and polished with a composite finishing and polishing kit

(Hawe Neos Dental) and silicone-carbide impregnated bristle brushes (Astrobrush,

Ivoclar Vivadent) in a slow-speed handpiece. Approximal finishing was performed

with flexible discs and abrasive strips (Soflex pop-on; 3M ESPE).

Fig. 1. X-ray preoparatory Fig. 2. Initial status

Fig. 3. The cavity preparation Fig. 4. IFPD made with fiber reinforced

composite.

86

Fig.5. The inlay bridges after the

cementation under rubber dam.

Fig. 6. Follow-up after 2 years.

Evaluation

Two independent examiners (C.M., M.C.) evaluated under magnification (Zeiss

3.6x35mm) all the restorations directly after the final polishing and after one week, 6,

12, 24 and 36 months. At baseline, and at 1-yr and 2-yr examinations, X-rays were

performed to check for surplus, marginal gaps, and secondary caries. During the

recalls, the patients came back to the office to answer questions about post-operative

sensitivity. Partial or total debonding of IFPDs, framework or resin composite

fractures and fiber exposures were considered failures. The restorations were

evaluated using the USPHS (Ryge et al 1973) modified parameters, to check their

stability and longevity in regard to the following characteristics: color match,

marginal discoloration, secondary caries, surface texture, marginal adaptation,

fracture and postoperative sensitivity. Alpha (α) means perfect condition, Bravo (β)

show restorations that are clinically acceptable, Charlie (c) indicate a need for

immediate replacement. In cases with only two decision possibilities, eg, debonding

or no debonding, the rating comprised only α or c (c= debonding). Plaque growth

and gingival health at the gingival pontic surfaces, abutment inlays and contralateral

control teeth were also measured using the plaque index (PI) and gingival index (GI).

Statistical analysis was applied to compare the restorations at the baseline and

after the last recall, and to check for differences between Group A and B. The

Wilcoxon’s matched-pairs signed ranks test measured the restorations’ success at the

87

appropriate time intervals and was used to rate all parameters. The Mann-Whitney

test was used to compare the data between the two groups. The null hypothesis was

rejected at the 5% level. The survival estimation method of Kaplan-Maier was used

with statistical software (JMP 5.1).

Table 2. Clinical results based on USPHS criteria after two-year.

Abutments

(n=78)

One week

α

Last

Recall α

One week

β

Last

Recall

β

One week

c

Last

Recall

c

Groups A B A B A B A B A B A B

Marginal

discoloration

38 40 37 40 - - - - - - 1 -

Secondary

caries

38 40 38 40 - - - - - - - -

Marginal

adaptation

38 40 34 40 - - 4 - - - - -

Postoperative

sensibility

23 38 36 40 5 2 - - 10 - 2 -

Debonding 38 40 35 40 - - - - - - 3 -

IFPDs (n=39)

Fracture

(including

chipping)

19 20 17 18 - - - - - - 2* 2

Surface

texture

18 20 18 20 1 - 1 - - - - -

Color match 17 17 17 16 2 3 2 4 - - - -

Fiber

exposure

19 20 19 18 - - - - - - - 2

* The micro fractures were visible only under SEM examination after the detachment of the

bridges.

88

RESULTS

The results are summarized in table2. During the observation time, two debondings

after 2 and 8 months were detected for the IFPDs luted with Excite DSC bonding

system. The detected bridges were immediately replaced with another inlay bridge

luted with Syntac. Some micro cracks in the pontic area of the two detached bridges

were observed under SEM (Fig. 7a-b). Two fiber exposures were noted after 8

months on the occlusal surface of one IFPD. The framework was visible under low

magnification (Zeiss 3.6x35mm) for self-evident color change. This kind of failure

was most probably due to the previous occlusal adjustment done after the

cementation that had left a thin layer of composite upon the fiber framework. The

bridge wasn’t replaced and is under observation now. Some hairline fracture of the

veneering materials near the connection between the pontic and the abutment were

detected in two cases (Fig. 8-9; a-b). No statistical differences regarding

detachment were showed between baseline and last recall between Group A and B

(P>0.05).

There was no fracture of the pontics, inlay retainers or inlay margins. The

IFPDs was always rated “alpha” with respect to secondary caries and surface

texture. The two debonded bridges showed “charlie” rating for the marginal

adaptation on six abutments and marginal discoloration on one abutment at the last

recall. Moderate to severe postoperative sensitivity was rated during the first six

months of the observation period. All hypersensitive teeth belonged to group A

luted with two-step bonding system. The higher values of sensitivity recorded

during the temporary restoration time were reported for group A. These patients

primarily reported strong chewing pain during the mastication and less temperature

sensitivity (10a-b). The dentinal sensitivity diminished after approximately 12

weeks and completely disappeared after six months in all cases except one

abutment that was endodontically treated in the retainer area without removal of

the bridge. The statistical analysis showed significant differences (P<0.05) for the

sensitivity between group A and group B at baseline and inside of group A during

the observation period. Group A showed 42.2% of dentinal sensitivity after 1

month unlike group B that recorded 0%. The percentage of “bravo” for the color

match was stable at 86.8% during the observation period and didn’t change at any

89

time. Plaque growth was moderate but no statistical differences occurred in PI and

GI between the abutments, pontics and contralateral teeth. The Kaplan-Maier

survival estimation was 89.4% for group A and 100% for the group B after 24

months. (Fig 11).

Fig. 7A. Gingival side of the

detached bridge. Some fractured

glass fibers are visible.

Fig. 7B. The pontic fibers are clearly visible

under magnification (x1000) after the

fracture of veneering composite.

Fig. 8A. The presence of old

restorations on the abutments near the

edentulous space is the typical situation

for the IFPDs.

Fig. 8B. The clinical condition doesn’t

permit to make a fiber framework with

ideal dimensions. This characteristic

could cause flexibility of the bridge.

Fig. 9A. The IFPD after 8 months Fig. 9B. The same bridge after 2 years.

The finger indicates the hairline fracture

of the veneering material in the pontic

element near the connection with the

inlay.

Fig. 10A. This image represents

impression of the dentin surface after the

hybridization with Excite DSC. Some

areas show the presence of hallows due

to the intrapulpal pressure that could

cause the hypersensitivity during the

chewing.

Fig. 10B. At 1000 magnification are

clearly visible the ditching created by

the dentinal fluid.

90

0,0

0,1

0,2

0,3

0,4

0,5

0,6

0,7

0,8

0,9

1,0S

urvi

ving

0 5 10 15 20 25Time (months)

Fig. 11. Kaplan-Meier estimation of the cumulative survival rate of IFPDs luted with

Excite DSC (group A) and Syntac (group B)

DISCUSSION

The results of this study demonstrate that IFPDs made with a light and heat–

polymerized microfilled composite bonded with FRC composite exhibited a high

percentage of clinical survival in a short observation period particularly in

combination with a classic 3-step adhesive bonding system but different discussion

points must be argued. The two important points to achieve success with this

conservative technique are the design of the fiber framework and the cavity

preparation. The design of the FRC framework and the position of the fibers can play

an important role in supporting the layering material in a continuous manner and to

strengthen the fiber sub-structure. Ellakwa (Ellakwa et al 2004) showed that the

different techniques of laboratory construction of fiber framework in the pontic area

significantly affected the fracture resistance of fiber-reinforced bridges. Maximizing

Group A

Group B

91

fiber volume fraction by increasing the proportion of fiber to composite should

significantly improve strength (Butterworth et al 2003). The position of the FRC

layer can have an effect on the flexural strength. Continuous unidirectional fibers

gave the highest strength and stiffness only in the direction of the fiber, while woven

fibers were able to reinforce the denture base polymers in two directions (De Boer et

al 1984).

The laboratory technique used in this study allows the extension in the vestibular and

buccal side to support the layering material like in metal framework and hold it in all

loading directions. The framework in the FRC was made in an anatomical shape in

the pontic element with parallel and weave fiberglasses. Vestibular and buccal FRC

extensions in the pontic element can increase the bonding area between the

framework and resin composite as well as hold the veneer material better during

occlusal loading. This observation suggested that modified design of frameworks

could increase the bond strength of resin composite to the fiber framework when

chewing (Monaco et al 2003). A similar approach was used by Freilich (Freilich et al

2002) that hypothesized the increased rigidity and a broader base of support provided

by the FRC substructure was needed to support the composite veneer. Thus, they

added a substantial amount of FRC bulk to the pontic component of the substructure

(low-volume design), resulting in the creation of the “high-volume” substructure

design and examined their relationship with the clinical performance. They observed

a 95 percent survival rate for the high-volume prostheses in contrast to a 62 percent

survival rate for the low-volume prostheses over a 3-year observation period. Similar

clinical results were shown in our previous study (Monaco et al 2003) on the

relationship between composition and substructure design and clinical performance

of Targis/Vectris system after a 4-year follow up period. The authors achieved a

97.5% survival rate for the framework built with parallel and woven fibers modifying

the design of the pontic element in an anatomical shape versus a 84% survival rate

for the restorations with simplified frameworks made only with parallel fibers.

Göhring (Göhring et al 2005) observed 36 posterior FRC IFPDs and after 5 years

they reported 71% of survival rate. Most failures were related to delaminations of the

veneering material and they are in agreement that modified framework design

92

significantly reduced the delamination rate (Göhring et al 2003). However, long-term

clinical studies must be conducted to confirm this hypothesis.

The detachments occurred in two clinical border-line cases. The first happened after

2 months in a patient with parafunctional habits whereas the second occurred after 8

months in a case with long span replacement (11 mm). The examination of the failed

bridges under SEM disclosed micro cracks in the gingival areas between the pontic

element and the inlay probably due to fatigue phenomena of the veneering material.

Lassila (Lassila et al 2004) achieved the highest flexural strength when the FRC layer

was located at the tension side of the test specimens. The particulate filler composite

is the weaker phase of the system. They demonstrated that when it is located on the

tension side, the fracture could easily initiate. The FRC structure benefits most when

the tensile stresses can be transferred to the reinforcing fibers. The veneering

particulate filler composite is strong in compression stress and, therefore, the FRC

structure requires less reinforcement fibers on the compression side. This area

represents a tensile zone and the SEM pictures showed the inner glass fiber of the

substructures.

Key elements for the clinicians include tooth preparation design that must allow an

adequate amount of FRC, an accurate interocclusal registration, and proper insertion

technique. An inadequate interocclusal registration could results in a considerable

occlusal adjustment by the dentist and the potential for inadvertent occlusal exposure

of the FRC substructure of an extremely thin layer of the composite veneer in

functional areas.

Color match of veneering composite (SR Adoro) was extremely stable unlike that of

the predecessor material (Targis) that showed in a previous study (Monaco et al

2003) a percentage of bravo in colour match of 29% at the last recall and the

deterioration was significant compared to the initial status. In this study, the clinical

evaluation of the FRC inlay bridges showed that the microfilled composite veneer

material exhibited good colour stability and resistance to wear. The surface texture

exhibited no change except for one case with a small chipped area most probably due

to a fabrication error.

Group A luted with Excite DSC showed a dentinal sensitivity slightly below 50%

after 1 month unlike group B luted with Syntac that didn’t record post-operative

93

sensitivity. Excite DSC is an ethanol-based, two step, dual-curable, single-bottle

adhesive. Total etch or simplified adhesives are more sensitive to the technique

because optimal hybridization and sealing of dentinal tubules with the wet bonding

technique may differ with each bonding system (Frankenberger et al 2000). Ferrari

(Ferrari et al 2003) confirmed the sensitivity of the technique with this adhesive

system showing in vivo no hybrid layer and extensive nanoleakage after excessive

drying and water tree formation along resin-dentin interfaces during excessive

wetting. Because the volatile adhesive solvent evaporates quickly, the continuous

transudation of dentinal fluid through open dentinal tubules before polymerization of

the adhesive may result in the entrapment of water-filled blisters along the adhesive

interface (Pashley et al 2002, Tay et al 2003). As the patient masticates, these blisters

may create a pumping effect that causes rapid movement of fluid through the tubules,

which in turn may trigger the A-delta nerve fibres in the pulpal-dentin complex

(Brännström et al 1972). Although most bonded restorations are retained because

there is a sufficient well-bonded surface area, a common clinical manifestation of

inconsistent bonding within a restoration is the patient’s complaint concerning post-

operative sensitivity (Unemori et al 2001). Clinically, no postoperative sensibility

was reported by the patients of the group B. This favorable outcome may be related

to the three-step adhesive bonding system associated to the methods used in the study

to seal the dentin before taking the impression. An accurate control times of the

primer and of the bonding agent (>20 s) can ensure against the postoperative

sensibility during the temporary period and after the final luting procedure.

Conclusion

In a short-term observation period, the FRC showed good clinical service but these

results need more observation in order to create a sound basis for the final

assessment of this restoration technique. The clinicians and the dental technicians

need to strictly follow indications, contraindications and instructions to achieve a

satisfactory clinical result. If a conservative inlay FPD is clinically indicated, the

patient must be informed that loss of sound hard tissue is minimal but durability of

a conventional ceramic fused to metal full coverage FPD and implant treatment is

more proven.

94

References

Behr M, Rosentritt M, Handel G. Fiber-reinforced composite crowns and FPDs: a

clinical report. Int J Prosthodont. 2003 May-Jun;16(3):239-43.

Behr M, Rosentritt M, Leibrock A, Schneider-Feyrer S, Handel G. In-vitro study of

fracture strength and marginal adaptation of fibre-reinforced adhesive fixed partial

inlay dentures. J Dent. 1999;27:163-168.

Beuchat M, Krejci I, Lutz F. Minimally invasive unreinforced adhesive

composite bridges: the clinical procedure. Schweiz Monatsschr Zahnmed.

1999;109:507-519.

Brännström M, Aström A. The hydrodynamics of the dentine: its possible

relationship to dental pain. Int Dent J 1972;22:219-227

Butterworth C, Ellakwa AE, Shortall A. Fibre-reinforced composites in

restorative dentistry. Dent Update. 2003;30:300-6.

DeBoer J, Vermilyea SG, Brady RE. The effect of carbon fiber orientation on

the fatigue resistance and bending properties of two denture resins. J Prosthet Dent

1984;51:119-121.

Ellakwa AE, Shortall AC, Marquis PM. Influence of different techniques of

laboratory construction on the fracture resistance of fiber-reinforced composite

(FRC) bridges. J Contemp Dent Pract 2004;4:001-013.

Ferrari M, Tay FR. Technique sensitivity in bonding to vital, acid-etched dentin.

Oper Dent. 2003 Jan-Feb;28:3-8.

Fischer H, Weber M, Marx R. Lifetime prediction of all-ceramic bridges by

computational methods. J Dent Res. 2003;238-242.

Frankenberger R, Kramer N, Petschelt A. Technique sensitivity of dentin

bonding: effect of application mistakes on bond strength and marginal adaptation.

Oper Dent 2000;25:324-330.

Freilich MA, Karmaker AC, Burstone CJ, Goldberg AJ. Development and

clinical applications of a light-polymerized fiber-reinforced composite. J Prosthet

Dent. 1998;80:311-318.

Freilich MA, Meiers JC, Duncan JP, Eckrote KA, Goldberg AJ. Clinical

evaluation of fiber-reinforced fixed bridges. J Am Dent Assoc. 2002

95

Nov;133(11):1524-34; quiz 1540-1.

Freilich MA, Meiers JC, Duncan JP, Eckrote KA, Goldberg AJ. Clinical

evaluation of fiber-reinforced fixed bridges. J Am Dent Assoc. 2002;133:1524-1534

Göhring TN, Mörmann WH, Lutz F. Clinical and scanning electron microscopic

evaluation of fiber-reinforced inlay fixed partial dentures: preliminary results after

one year. J Prosthet Dent 1999; 82: 369-377.

Göhring TN, Roos M. Inlay-fixed partial dentures adhesively retained and reinforced

by glass fibers: clinical and scanning electron microscopy analysis after five years.

Eur J Oral Sci, 2005; 113: 60–69.

Göhring TN, Schmidlin PR, Lutz F. Two-year clinical and SEM evaluation of glass-

fiber-reinforced inlay fixed partial dentures. Am J Dent. 2002;15:35-40.

Göhring TN, Zappini G, Mayer J, Zehnder M. Optimization of glass-fiber

framework for fixed partial dentures: Laser-interferometrical analysis. Quintessenz

Int 2003; 35: 668–675.

Gruica B, Wang HY, Lang NP, Buser D. Impact of IL-1 genotype and smoking

status on the prognosis of osseointegrated implants. Clin Oral Implants Res.

2004;15:393-400.

Kern M, Knode H, Strubb JR. The all-porcelain, resin-bonded bridge.

Quintessence Int. 1991;22:257-262.

Kern M, Strub JR, Lu XY. Wear of composite resin veneering materials in a

dual-axis chewing simulator. J Oral Rehabil. 1999;26:372-378.

Knobloch LA, Kerby RE, Seghi R, van Putten M. Two-body wear resistance and

degree of conversion of laboratory-processed composite materials. Int J Prosthodont.

1999;12:432-438.

Kolbeck C, Rosentritt M, Behr M, Lang R, Handel G. In vitro examination of

the fracture strength of 3 different fiber-reinforced composite and 1 all-ceramic

posterior inlay fixed partial denture systems. J Prosthodont. 2002;11:248-253.

Lassila LVJ, Vallittu PK. The effect of fiber position and polymerization

condition on the flexural properties of fiber-reinforced composite. J Contemp Dent

Pract 2004;2:014-026.

Loose M, Rosentritt M, Leibrock A, Behr M, Handel G. In vitro study of

fracture strength and marginal adaptation of fiber-reinforced-composite versus all

96

ceramic fixed partial dentures. Eur J Prosthodont Restor Dent. 1998;6:55-62.

Meiers JC, Freilich MA. Chairside prefabricated fiber-reinforced resin

composite fixed partial dentures. Quintessence Int. 2001;32:99-104.

Monaco C, Ferrari F, DDS, Miceli GP, Scotti R. Clinical Evaluation of Fiber-

Reinforced Composite Inlay Fixed Partial Dentures. Int J Prosthodont 2003;16:319-

325.

Monaco C, Miceli GP, Scotti R. Die mit dem neuen, mikrogefüllten Komposit-

Material SR Adoro verblendete Inlay-Brücke. Quintessenz Zahntech 2003;3:292-

305.

Morin D, DeLong R, Douglas WH. Cusp reinforcement by the acid-etch

technique. J Dent Res. 1984;63:1075-1078

Pahley DH, Pashley EL, Carvalho RM, Tay FR. The effects of dentin

permeability on restorative dentistry. Dent Clin North Am 2002;46.211-245.

Quirynen M, De Soete M, van Steenberghe D.Infectious risks for oral implants:

a review of the literature. Clin Oral Implants Res. 2002;13:1-19.

Report on base metal alloy for crown and bridge applications: benefits and risks.

Council on Dental Materials, Instruments and Equipment. J Am Dent Assoc

1995;111:479-483.

Ryge G, Snyder M. Evaluating the clinical quality of restorations. J Am Dent

Assoc 1973;87:369-377.

Sewon LA, Ampula L, Vallittu PK. Rehabilitation of a periodontal patient with

rapidly progressing marginal alveolar bone loss: 1-year follow-up. J Clin

Periodontol. 2000;27:615-619.

Tay FR, Pashley DH. Have dentin adhesives become too hydrophilic? J Can

Dent Assoc 2003;69:726-731.

Unemori M, Matsuya Y, Akashi A, Goto Y, Akamine A. Composite resin restoration

and postoperative sensitivity: clinical follow-up in an undergraduate program. J Dent

2001;29:7-13.

Vallittu PK. Survival rates of resin-bonded, glass fiber-reinforced composite

fixed partial dentures with a mean follow-up of 42 months: a pilot study.J Prosthet

Dent. 2004;91:241-246.

Zumbuhl R, Lutz F, Krejci I. In-vitro studies of resin-bonded slot composite bridges

97

compared to conventionally prepared composite bridges. Schweiz Monatsschr

Zahnmed. 2000;110:505-522.

98

Chapter 9: Alternative materials as regards FRC

Fatigue test in shear: its effect on bond of a glass-infiltrated alumina ceramic to

human dentin, using different luting procedures.

99

The concepts of minimally invasive dentistry combined with adhesive technique

have been highly recommended because it’s based on the conservation of dental

structure. (Degrange et al 1997). Hence, the ceramic partial restorations compared

with complete crowns present this conservative approach because they are minimally

invasive. Thus, it is important that the enamel/dentin and the ceramic substrate are

appropriately conditioned to optimize this adhesive process. (Özcan 2002, Özcan &

et al 2002, Özcan 2003, Özok et al 2004,Reis et al 2004).

With the increase of aluminum oxide content (Al2O3) in feldspar ceramics,

there is a significant improvement in the mechanical properties allowing metal-free

restorations to be employed more predictably (Tinschert et al 2000, Guazzato et al

2004). In posterior teeth with high mechanical loading, the ceramics with high

crystalline content (aluminum and/or zirconium oxides) have presented better clinical

results than feldspar-, leucite-, and lithium disilicate-based ceramics (Scotti et al

1995, Oden et al 1998, Hayashi et al 2000, Fradeani et al 2002). However the

increase of the mechanical strength, by increasing the crystalline content and

decreasing the glass content, results in an acid-resistant ceramic whereby any type of

acid conditioning produces insufficient surface changes for adequate bonding to

resin. (Awliya et al., 2000, Derand et al 2000, Madani et al 2000, Ozcan et al 2001,

Della Bona et al 2002, Özcan et al 2003).

Thus, even though the etching with hydrofluoric acid combined with the

silanization is the principal conditioning method of feldspar-, leucite- and lithium

disilicate-based ceramics (silica-based ceramic) (Della Bona et al 2000, Debnath et al

2003, Borges et al 2003, Özcan et al 2003, Della Bona 2004, Melo et al 2005). This

conditioning method does not allow high bond strength to acid-resistant ceramics

(Awliya et al 2000, Derand et al 2000, Madani et al 2000, Ozcan et al 2001, Della

Bona et al 2002). Currently, the tribochemical silica coating method appears to be the

choice conditioning for the acid-resistant ceramic. (Kern et al 1995, Awliya et al

1998, Özcan et al, 2001, Özcan et al 2003, Bottino et al 2005, Valandro et al 2005,

Valandro et al 2005, Amaral et al 2005). Due to the high-speed surface impact of the

alumina particles modified by silica, there is an embedding of these particles on the

ceramic surface (Sun et al 2000). Consequently, there is a chemical bond between the

silica coated ceramic surface and the silane agent (y-

methacryloxypropyltrimethoxysilane), and the latter to the resin material.

100

Furthermore these tribological concepts, the resin cements with MDP

monomers (10-methacryloyloxydecyl-dihydrogen-phosphate) have allowed high

bond strength to acid-resistant ceramics, as well as high bonding durability, due to

the bond between the phosphate monomers and the aluminum/zirconium oxides of

the dental ceramic (Kern et al 1998, Wegner et al 2000, Özcan et al 2001, Friederich

et al 2002, Wegner et al 2002, Blatz et al 2003 Özcan et al 2003, Blatz et al 2004,

Hummel et al 2004, Leite et al 2005).

Bond strength- (Ozturk et al 2003) and microleakage-tests (Jacques et al

2003) can be used to evaluate the bonding between the dental substrate and the

ceramic. However, mechanical fatigue tests conducted in a humid environment seem

to be the best in vitro method to predict the clinical performance of dental materials

and restorative techniques. Fatigue is a type of test that may lead to the fracture of a

structure after repeated loading, and may be explained by the spread of microscopic

cracks from areas of force concentration, usually in areas containing macroscopic or

molecular structural defects (Pontius et al 2002, Scotti et al 2002, Wiskott et al

1995). Pontius (Pontius et al 2002) described that 1,200,000 cycles in a mechanical

fatigue test correspond to approximately 5 years of clinical function.

Thus, the purpose of this study was to evaluate the fatigue resistance of the

bonding between human dentin and glass-infiltrated aluminum-based ceramic, using

different luting materials. The hypothesis was that the fatigue resistance can be

modified depending on used luting procedures.

MATERIAL AND METHODS

Selection and preparation of teeth

Forty extracted, caries-free human third molars were used. The teeth were

collected in the Dept. of Oral Science (University of Bologna), after obtaining the

patient’s informed consent. After the extraction, the teeth were immediately

immersed in 0.02%-thimol/24h and stored in distilled water. Part of the crown

was removed, exposing a flat dentin surface of approximately 3mm x 4mm,

which was polished on wet #320, 600, 1000-grit silicon carbide (SiC) paper,

respectively, for 60s to standardize the smear layer.

101

After preparation, each specimen was embedded in a PVC cylinder

(height: 20mm, diameter: 10mm) filled with chemically cured epoxy resin

(Epoxy Resin 285, Schaller, Florence, Italy), remaining exposed only the dentin

surface for luting.

Preparation of alumina ceramic samples

Forty ceramic blocks of a glass-infiltrated alumina ceramic (In-Ceram Alumina

for Cerec 2) (VITA Zahnfabrik, Bad Säckingen, Germany) were fabricated

according to the manufacturer's specifications. These blocks were 3mm x 3mm x

4mm in dimension and the bonding surfaces (3mm x 3mm) of each block were

polished using a 600- to 1200-grit metallographic (3M, St. Paul, USA) in a

polishing machine. The surfaces of all the ceramic blocks were submitted to the

chairside airborne abrasion with 110µm aluminum oxide particles (Al2O3),

following this blasting protocol: a = perpendicular to the surface; b = 10-mm

distance; c = 20-second time; d = 2.8-bar pressure.

Luting procedures

The materials used in this study and the luting technical procedures are described

in Table 1. Four luting protocols were tested following the manufacturer’s

recommendations (luting system [LS] / ceramic surface conditioning [CSC]). The

tooth- and ceramic- sp were randomly divided in four groups (n=10):

GROUP 1: [LS] RelyX Unicem / [CSC] Al2O3;

GROUP 2: [LS] One-Step + Duo-Link / [CSC] Al2O3 + etching with 4%

hydrofluoric acid (Porcelain Etchant) + silane coupling agent (Porcelain Primer).

GROUP 3: [LS] Panavia F / [CSC] Al2O3.

GROUP 4: [LS] Scotchbond 1 + RelyX ARC (3M-Espe) / [CSC] Al2O3 + chairside

airborne abrasion with siliceous acid modified by 30µm Al2O3 particles (Si2Ox)

(blasting procedures: identical to the airborne abrasion with Al2O3) + silane

coupling agent (Tribochemical silica coating) (CoJet System);

102

After the ceramic conditioning and application of the adhesive systems

(excepted in G1), the luting cements were manipulated as recommended by the

manufacturers and applied on the bonding surface of each ceramic block, which

was than cemented to dentin surface. The set-up remained under the constant load

of 750g for 10min and during this period the cement excesses were removed. The

specimens (sp) were stored in distilled water to 37oC for 24 h and then were

submitted to the fatigue test.

Mechanical fatigue test in shear (Figure 2 and 3)

Initially, all sp were grasped in an apparatus, which allows the dentin surface to

be almost perpendicular to the fatigue test. A metallic cylinder were then

positioned on the ceramic surface (perpendicular to dentin surface) and the sp

were subjected to 1 million fatigues cycling shear with sinusoidal load ranging

from 0 to 21N at 8Hz frequency and 37°C water irrigation.

Fatigue resistance scores and statistical analysis

After the fatigue test, it was given a classification (score) for fractured and no-

fractured sp based on the number of the mechanical cycle:

Score 0: samples fractured before cycling.

Score 1: samples fractured between 0 and 199,999 cycles.

Score 2: samples fractured between 200,000 and 399,999 cycles.

Score 3: samples fractured between 400,000 and 599,999 cycles.

Score 4: samples fractured between 600,000 and 799,999 cycles.

Score 5: samples fractured between 800,000 and 999,999 cycles.

Score 6: samples non-fractured up to 1 million cycles.

The collected data were submitted to statistical analysis (Kruskal-

Wallis test and post hoc Tukey test), using a level of significance of 5%.

103

Analysis of the failure modes

The fractured sp during the fatigue test were analyzed in a Scanning Electron

Microscopy (magnification from x200 to x1000) (JEOL–JSM–5400, Jeol Ltd,

Tokyo, Japan) in order to evaluate the sp failure mode, which were classified in

seven types: Type 1) fracture of the luting cement (cohesive); Type 2) fracture of

the dentin (cohesive); Type 3) cohesive fracture of the ceramic (cohesive); Type

4) failure between dentin and adhesive system (adhesive failure); Type 5) failure

between luting cement and ceramic (adhesive failure); Type 6) mixed failure

(type 1 + type 4); Type 7) mixed failure (type 1 + type 5); Type 8) mixed failure

(type 4 + adhesive failure between resin cement and bonding agent). All

micrographics were analyzed by three calibrated observers.

RESULTS

The Kruskal-Wallis nonparametric analysis of the fatigue resistance scores are

shown in Table 2, in which it was possible to note that there were differences

among the studied groups.

The mean score values and standard deviations of the groups are

shown in Table 3, noting that G3 (Score = 5.9 [1 failure]) and G4 (Score = 6

[none failure]) were statistically similar (P=0.33) and had significantly the

highest fatigue resistance, when compared to G1 (Score = 3.9 [5 failures]) and G2

(Score = 3.7 [6 failures]) (P<.03).

In the fractured sp microscopic analyses it was observed that: (1) G1 –

5 fractures type 4 (Figure 4); (2) G2 – 6 fractures type 5 (Figure 5); G3 – 1

fracture type 8 (Figure 6).

DISCUSSION

The hypothesis of this study was confirmed, since the fatigue resistance was

different depending on the protocol used to lute the glass infiltrated alumina

104

ceramic to the dentin. The highest fatigue resistances were obtained whether by

the combination of self-etching primer with MDP-based resin cement (G3) or the

combination of total etching/single bottle with resin cement/tribochemical silica

coating (G4).

The results for G3 may be explained by the combination of the

chemical bonds between the MDP phosphate monomers and the aluminum oxides

(Wegner et al 2000, Özcan et al 2001, Blatz et al., 2003, Hummel et al 2004,

Leite et al 2005), and the dentin treated with self-etching primer. (Uno et al 2000,

Varela et al 2003) For G4, the results may be explicated by the chemical bond

among silica coated ceramic surface, silane agent and the resin material, (Kern et

al 1995), combined with the adequate bonding of the adhesive system to the

dentin (Abu-Hanna et al 2004, Helvatjoglu-Antoniades et al 2004, Reis et al

2004).

The performance of resin cements that contain MDP phosphate

monomers (10-methacryloyloxydecyl-dihydrogen-phosphate) (G3) have been

observed by various studies, which noted a high bonding durability of this cement

to acid-resistant ceramics, such as glass-infiltrated alumina/zirconium ceramic

(Kern et al 1995, Ozcan et al 2001), yttrium-oxide-partially-stabilized zirconia

ceramic (Kern et al 1998, Wegner et al 2000, Wegner et al 2002, Blatz et al

2004), densely-sintered alumina ceramic (Friederich et al 2002, Blatz et al. 2003,

Hummel et al 2004).

Despite the resin cement of G3 was bonded to the acid-resistant

ceramic, one sp showed a failure type 8, since the fracture occurred in two zones:

partial fracture between dentin and adhesive system, and partial fracture between

the adhesive system and resin cement (Panavia). Even though this fracture has not

been statistically significant, since G3 (1 fracture) was statistically similar to G4

(none fracture), it maybe explained due to the features of the dentin adhesive

system and the resin cement employed in G3. The employed self-etching primer

adhesive system exhibited a low pH value, producing a chemical match with the

polymerization chemical process of the dual-cure resin cement. Some studies

have indicated a possible chemical incompatibility between adhesive systems

105

with low pH and resinous materials of chemical- and dual-polymerization.

(Sanares et al 2001, Cheong et al 2003, Suh et al 2003, Tay et al 2003, Tay et al

2003). A correlation was observed between the decline in microtensile bond

strengths of chemical-cured composites coupled to bonded dentin and the acidity

of these adhesives. (Sanares et al 2001) Since single-step self etch adhesives

contain a higher concentration of acidic resin monomers, the coupling of

chemical/dual-cured composites to the hydrated dentin bonded with these

adhesives were also found to be inferior to those achieved by using light-cured

composites. It is known that acidic resin monomers retard the polymerization of

chemical/dual-cured composites that are initiated via peroxide-amine type binary

redox catalysts. Interaction between acidic adhesive resin monomers and the

basic composite tertiary amines results in the consumption of the latter in acid-

base reactions, depriving their capacity of generating free radicals in subsequent

redox reactions. (Suh et al 2003, Tay et al 2003).

In this current study, it was observed that none of the G4 specimens

fractured during the fatigue test, meaning that the two interfaces (resin cement–

ceramic; dentin-adhesive system) supported the fatigue. The high bond-strength

and -durability promoted by tribochemical silica coating combined with resin

cement has also been stated by studies. (Kern et al 1995, Awliya et al 1998,

Özcan et al 2001, Özcan et al 2003, Bottino et al 2005, Valandro et al 2005,

Valandro et al 2005). The fatigue resistance of the interface cement-ceramic can

be explained by the following phenomena: the airborne particle abrasion allows

the ceramic surface to be embedded by silica oxide coated aluminum particle

(Sun et al 2000) forming the silica-modified surface chemically more reactive to

the resin via silane coupling agents. Silane molecules react with water to form

three silanol groups (–Si–OH) from the corresponding methoxy groups (–Si–O–

CH3). (Plueddemann 1970, Ozcan et al 2003, Ozcan et al 2004) The silanol

groups then react further to form a siloxane (–Si–O–Si–O–) network with the

silica surface. Monomeric ends of the silane molecules react with the

methacrylate groups of the adhesive resins in a free radical polymerization

process. (Plueddemann 1970, Jedynakiewicz et al 2001, Ozcan et al 2003, Ozcan

et al 2004). Summarizing, the employed tribochemically assisted system

106

(chairside system) allows a chemical bond among the coated silica, silane

coupling agent and resin cement. Regarding the fatigue resistance of the interface

dentin-adhesive system, some studies stated that the adhesive system used in this

group (G4) present high dentin bond strength (Abu-Hanna et al 2004,

Helvatjoglu-Antoniades et al 2004, Reis et al 2004), due to the hybridization of

dentin (Koshiro et al 2004), as well as the compatibility between the bonding

agent and the resin cement (Sanares et al 2001), which justified the resistance to

the fatigue.

Analyzing the micrographics of the 5 fractured sp of G1 (Figure 4), it

was noted that the failure occurred in the interface dentin – resin cement (type 4).

In this micrographic, the dentin surface coated by the smear layer can also be

observed. This smear layer was not removed because the dentin is not etched and

any bonding agent was applied as recommended by the manufacturer. Thus the

hybrid layer is not formed and the bonding between the cement and the dentin did

not support the mechanical fatigue, explaining the type of fracture.

In G2, 6 fractures type 5 were observed (failures between the resin

cement and ceramic surface). The ceramic surface conditioning (4% hydrofluoric

acid + silane coupling agent) recommended for this luting protocol justified the

noted fracture pattern. Currently, it is stated that the hydrofluoric acid does not

present ability to attack the compact ceramic surface with high alumina content

and low vitreous phase (acid-resistant ceramic), such as the ceramic employed in

the current study (glass-infiltrated alumina ceramic) (Awliya et al 1998, Derand

et al 2000, Madani et al 2000, Ozcan et al 2001, Della Bona et al 2002 Borges et

al 2003, Özcan et al 2003). This glass infiltrated high-alumina core ceramic is

composed of approximately 80w% of crystalline phase (aluminum oxide) and

20w% of glassy phase (aluminum-silicate-lanthanum) (Tinschert et al 2000,

Guazzato et al 2004). Hence, due to the low silica content in this ceramic, the

application of the silane agent also does not contribute to bonding, because the

MPS-based silane agent presents a great chemical bond to silica oxides and a

weak bond to aluminum oxides, which is the basis ingredient of alumina ceramic

(Plueddemann 1970, Kern et al 1994, Ozcan et al 2003, Ozcan et al 2004).

Exactly as stated by Valandro (Valandro et al 2005) the ceramics can be

107

classified into groups depending on how the surface is attacked by the

hydrofluoric acid (HF). Acid-sensitive ceramics are promptly etched by HF

resulting in micro-mechanical retentive ceramic surfaces (e.g.: feldspar-, leucite-

and lithium disilicate-based ceramics) (Della Bona et al 2000, Debnath et al 2003,

Borges et al 2003, Özcan et al 2003, Della Bona et al 2004, Melo et al 2005).

Acid-resistant ceramics do not suffer much surface degradation by HF etching,

preventing a reliable micro-mechanical bond to resin (e.g.: glass-infiltrated

alumina and zirconia ceramic systems, densely-sintered alumina ceramics, and

yttrium-oxide-partially-stabilized zirconia ceramics) (Ozcan et al 2003, Bottino et

al 2005, Valandro et al 2005, Amaral et al 2005).

However, considering the micrographics of the fractured sp of G2, it

can be noted that the interface dentin – adhesive system resisted to the fatigue

test. Thus, the bonding of the G2 adhesive system to the dentin maybe considered

safe, in light of the parameters followed in this study. Some studies have shown

that One Step adhesive system presents high bond strength to the dentin (Cheong

et al 2003, Abdalla et al 2004).

Taking into account the results obtained in G2, in which the bonding to

the dentin was stable and the bonding to the ceramic did not resist to mechanical

fatigue, the clinicians should reflect on the choice of the adhesive system to bond

the dentin and the ceramic surface. For example, if the manufacturer’s

instructions regarding the ceramic surface conditioning were not followed,

namely if the surface ceramic had been conditioned with a tribochemical silica

coating method, in view of the current concepts of adhesion to acid-resistant

ceramic, theoretically the fractured observed fracture in the interface resin cement

– ceramic could not have occurred. The clinicians should attentively observe

some aspects during the luting of all-ceramic restorations. Even though the

manufacturer’s instructions should obviously be considered, a global view

regarding the current concepts to treat the dental hard tissues, as well as the

conditioning of the ceramic surface must be considered in order to optimize the

global performance of the bonded ceramic restorations.

108

Surely, other factors such as the case preview must be planned, the

characteristics of the prosthetic preparations and generally the conditions for the

cementation can affect the global performance, and therefore further clinical

evaluation studies are necessary to state the effect of the change of luting

protocols or the mixing of different protocols. The combination of dentin

adhesive systems with MDP-based resin cement or resin cement/tribochemical

silica coating were the best luting protocols for aluminous ceramic. The

cementation strategies of glass-infiltrated alumina ceramic recommended by the

manufacturers maybe eventually changed in order to optimize the bond to all the

substrate.

Abdalla AI. Microtensile and tensile bond strength of single-bottle adhesives: a

new test method J Oral Rehab 2004;31:379-84.

Abu-Hanna A, Gordan VV. Evaluation of etching time on dentin bond strength

using single bottle bonding systems J Adhes Dent 2004;6:105-10.

Amaral R, Leite FPP, Bottino MA, Valandro LF, Özcan M. Microtensile bond

strength of a resin cement to glass infiltrated zirconia-reinforced ceramic: the effect

of surface conditioning. Dent Mater 2005. (forthcoming)

Awliya W, Odén A, Yaman P, Dennison JB, Razzoog ME. Shear bond strength

of a resin cement to densely sintered high-purity alumina with various surface

conditions. Acta Odontol Scand 1998;56:9-13.

Blatz MB, Sadan A, Arch GH, Lang BR. In vitro evaluation of long-term

bonding of Procera AllCeram alumina restorations with a modified resin luting agent.

J Prosthet Dent 2003;89:381-387.

Blatz MB, Sadan A, Martin J, Lang BR. In vitro evaluation of shear bond

strengths of resin to densely-sintered high purity zirconium-oxide ceramic after long-

term storage and thermal. J Prosthet Dent 2004;91:356-362.

Borges GA, Sophr AM, Goes MF, Correr Sobrinho L, Chan DCN. Effect of

etching and airborne particle abrasion on the microstructure of different dental

ceramics. J Prosthet Dent 2003;89:479-488.

Bottino MA, Valandro LF, Scotti R, Buso L. Effect of surface treatments on the

109

resin bond to zirconium-based ceramic. Int J Prosthodont 2005;18:60-65.

Cheong C, King NM, Pashley DH, Ferrari M, Toledano M, Tay FR.

Incompatibility of self-etch adhesives with chemical/dual-cured composites: two-step

vs one-step systems. Oper Dent 2003;28:747-755.

Consani S, Santos JG, Correr Sobrinho L, Sinhoreti MA, Sousa-Neto MD. Effect

of cement types on the tensile strength of metallic crowns submitted to

thermocycling. Braz Dent J 2003;14(3):193-196.

Degrange M, Roulet JF. Minimally invasive restorations with bonding.

Germany: Quintessence, 1997.

Della Bona A, Anusavice KJ. Microstructure, composition, and etching

topography of dental ceramics. Int J Prosthodont 2002;15:159-167.

Dérand P, Dérand T. Bond strength of luting cements to zirconium oxide

ceramics. Int J Prosthodont 2000;13:131-135.

Dérand P, Dérand T. Bond strength of luting cements to zirconium oxide

ceramics. Int J Prosthodont 2000;13:131-135.

Fradeani M, Redemagni M. An 11-year clinical evaluation of leucite-reinforced

glass-ceramic crowns: a retrospective study. Quintessence Int. 2002;33:503-510.

Friederich R, Kern M. Resin bond strength to densely sintered alumina ceramic.

Int J Prosthodont 2002;15:333-338.

Gemalmaz D. Use of heat-pressed, leucite-reinforced ceramic on anterior and

posterior onlays: a clinical report. J Prosthet Dent 2002;87(2):133-135.

Guazzato M, Albakry M, Ringer SP, Swain MV. Strength, fracture toughness

and microstructure of a selection of all-ceramic materials. Part I. Pressable and

alumina glass-infiltrated ceramics. Dent Mater 2004;20:441-448.

Hayashi M, Tsuchitani Y, Kawamura Y, Miura M, Takeshige F, Ebisu S. Eight-

year clinical evaluation of fired ceramic inlays. Oper Dent. 2000;25:473-481.

Helvatjoglu-Antoniades M, Koliniotou-Kubia E, Dionyssopoulos P. The effect

of thermal cycling on the bovine dentine shear bond strength of current adhesive

systems. J Oral Rehab 2004;31:911-917.

Hummel M, Kern M. Durability of the resin bond strength to the alumina

ceramic Procera. Dent Mater 2004;20:498-508.

Jacques LB, Ferrari M, Cardoso PE. Microleakeage and resin cement film

110

thickness of luted all-ceramic and gold electroformed porcelain-fused-to-metal

crowns. J Adhes Dent 2003;5:145-152.

Jedynakiewicz NM, Martin N. The effect of surface coating on the bond strength

of machinable ceramics. Biomaterials 2001;22:749-52.

Kern M, Thompson VP. Bonding to glass infiltrated alumina ceramic: adhesive

methods and their durability. J Prosthet Dent 1995;73:240-249.

Kern M, Thompson VP. Bonding to glass infiltrated alumina ceramic: adhesive

methods and their durability. J Prosthet Dent 1995;73:240-249.

Kern M, Thompson VP. ESCA surface characterization of the alumina ceramic

In-Ceram after various conditioning methods for resin bonding. J Dent Res

1994;73(sp iss):197. (Abstract #763)

Kern M, Wegner SM. Bonding to zirconia ceramic: adhesion methods and their

durability. Dent Mater 1998;14:64-71.

Koshiro K, Inoue S, Tanaka T, Koase K, Fujita M, Hashimoto M, Sano H. In

vivo degradation of resin–dentin bonds produced by a self-etch vs. a total-etch

adhesive system. Eur J Oral Sci 2004; 112: 368-375.

Kramer N, Lohbauer U, Frankenberger R. Adhesive luting of indirect

restorations. Am J Dent 2000;13:60D-76D.

Leite FPP, Valandro LF, Guidini A, Bottino MA, Kimpara ET. Evaluation of the

tensile bond strength between an aluminous ceramic and two resin cements. Braz

Dent Sci 2005. (forthcoming)

Lopes GC, Baratieri CM, Baratieri LN, Monteiro S Jr, Cardoso Vieira LC.

Bonding to cervical sclerotic dentin: effect of acid etching time. J Adhes Dent

2004;6:19-23.

Madani M, Chu FCS, McDonald AV, Smales RJ. Effects of surface treatments

on shear bond strengths between a resin cement and an alumina core. J Prosthet Dent

2000;83:644-647.

Melo RM, Valandro LF, Leite F, Bottino MA. Bonding to a leucite reinforced

feldspar ceramic. Braz Dent J 2005. (in press)

Nakabayashi N, Pashley DH. Hybridization of dental hard tissues. Tokyo:

Quintessence, 2000.

Odén A, Andersson M, Krystek-Ondracek I, Magnusson D. Five-year clinical

111

evaluation of Procera AllCeram crowns. J Prosthet Dent 1998;80:450-456.

Özcan M, Akkaya A. New approach to bonding all-ceramic adhesive fixed

partial dentures: A clinical report. J Prosthet Dent 2002;88: 252-254.

Özcan M, Alkumru H, Gemalmaz D. The effect of surface treatment on the shear

bond strength of luting cement to a glass-infiltrated alumina ceramic. Int J

Prosthodont 2001;14:335-339

Özcan M, Matinlinna JP, Vallittu PK, Huysmans MC. Effect of drying time of 3-

methacryloxypropyltrimethoxysilane on the shear bond strength of a composite resin

to silica-coated base/noble alloys. Dent Mater 2004;20:586-590.

Özcan M, Vallittu PK. Effect of surface conditioning methods on the bond

strength of luting cement to ceramics. Dent Mater 2003;19:725-731.

Özcan M. Evaluation of alternative intra-oral repair techniques for fractured

ceramic-fused-to-metal restorations. J Oral Rehab 2003; 30:194-203.

Özcan M. The use of chairside silica coating for different dental applications: A

clinical report J Prosthet Dent 2002;87:469-472.

Özok AR, Wu M-K, Gee AJ, Wesselink PR. Effect of dentin perfusion on the

sealing ability and microtensile bond strengths of a total-etch versus an all-in-one

adhesive. Dent Mater 2004; 20, 479-486.

Ozturk N, Aykent F. Dentin bond strengths of two ceramic inlay systems after

cementation with three different techniques and one bonding system. J Prosthet Dent

2003;89:275-281.

Plueddemann EP. Adhesion through silane coupling agents. J Adhesion

1970;2:184-201.

Pontius O, Hutter JW. Survival rate and fracture strength of incisors restored

with different post and core systems and endodontically treated incisors without

coronoradicular reinforcement. J Endod 2002;28:710-715.

Reis A, Loguercio AD, Carvalho RM, Grande RHM. Durability of resin dentin

interfaces: effects of surface moisture and adhesive solvent component. Dent Mater

2004; 20:669-676.

Sanares AME, Itthagarun A, King NM, Tay FR, Pashley DH. Adverse surface

interactions between one-bottle light-cured adhesives and chemical-cured

composites. Dent Mater 2001;17:542-556.

112

Scotti R, Catapano S, D’Elia A. A clinical evaluation of In-Ceram crowns. Int J

Prosthodont 1995;8:320-323.

Scotti R, Ferrari M. Proprietà mechaniche e valutazione in vitro. In: Perni in

fibra: Presupposti teorici ed applicazioni cliniche. Milan: Masson 2002;39-51.

Suh BI, Feng L, Pashley DH, Tay FR. Factors contributing to the incompatibility

between simplified-step adhesives and chemically-cured or dual-cured composites.

Part III. Effect of acidic resin monomers. J Adhes Dent 2003;5:267-282.

Sun R, Suansuwan N, Kilpatrick N, Swain M. Characterization of

tribochemically assisted bonding of composite resin to porcelain and metal. J Dent

2000;28: 441-445.

Tay FR, Pashley DH, Yiu CK, Sanares AM, Wei SH. Factors contributing to the

incompatibility between simplified-step adhesives and chemically-cured or dual-

cured composites. Part I. Single-step self-etching adhesive. J Adhes Dent 2003;5:27-

40.

Tay FR, Pashley DH. Resin bonding to cervical sclerotic dentin: a review. J Dent

2004;32:173-196.

Tay FR, Suh BI, Pashley DH, Prati C, Chuang SF, Li F. Factors contributing to

the incompatibility between simplified-step adhesives and self-cured or dual-cured

composites. Part II. Single-bottle, total-etch adhesive. J Adhes Dent 2003;5:91-105.

Tinschert J, Zwez D, Marx R, Anusavice KJ. Structural reliability of alumina-,

feldspar-, leucite-, mica- and zirconia-based ceramics J Dent 2000;28:529-535.

Uno S, Tanaka T, Kawamoto C, Konishi J, Sano H Microtensile bond strength to

dentin and cavity adaptation of Cerec 2 inlay restoration Am J Dent 2000;13:59.

Valandro LF, Della Bona A, Bottino MA, Neisser MP. Bonding to densely

sintered alumina ceramic: The effect of ceramic surface treatment. J Prosthet Dent

2005;93:…-…..

Valandro LF, Leite FPP, Scotti R, Bottino MA, Neisser MP. Effect of ceramic

surface treatment on the microtensile bond strength between a resin cement and an

alumina-based ceramic. J Adhes Dent 2004; 6(4):327-332.

Varela SG, Rábade LB, Lombardero PR, Sixto JML, Bahillo JDG, Park SA. In

vitro of endodontic post cementation protocols that use resin cements. J Prosthet

Dent 2003;89:146-153.

113

Wegner SM, Gerdes W, Kern M. Effect of different artificial aging conditions

on ceramic-composite bond strength. Int J Prosthodont 2002;15:267-272.

Wegner SM, Kern M. Long-term resin bond strength to zirconia ceramic. J

Adhesive Dent 2000;2:139-147.

Wiskott HWA, Nicholls JL, Belser UC. Stress fatigue: Basic principles and

prosthodontic implications. Int J Prosthodont 1995;8:105-116.

.

114

Chapter 10 Other clinical application of FRC

Clinical evaluation of teeth restored with quartz fiber-reinforced epoxy resin posts.

115

Both researchers and manufacturers have introduced several post-and-core

restorations with the aim of providing reliable systems for reconstruction of

endodontically treated teeth. In spite of these efforts, it is still difficult to predict the

clinical survival times of treated teeth restored with posts and cores (Goodacre et al

1994).

The prognosis is related to several factors, including the type of material used for the

post and core; the shape, dimensions, and length of the post; and the kind of cement

used. The major disadvantage associated with conventional cast-metal posts is

vertical root fracture. Having high rigidity, metal posts appear to vibrate at high

frequencies when loaded with lateral forces. The focusing of these forces in

unpredictable “critical points” may determine longitudinal fractures of the root or

metal corrosion and consequently lead to loss of the tooth (Trabert et al 1984,

Morgano et al 1993, Duret et al 1990).

In 1990, Duret (Duret et al 1990) proposed carbon-fiber posts, among the many

prefabricated fiber post-and-core systems, to reduce the failure rate. These relatively

recent posts are made of equally aligned carbon fibers attached to an epoxy resin

matrix and present an interesting property, anisotropic behaviour. In other words, the

material has different physical responses when loaded in different directions. This

characteristic is of clinical relevance, as it may strongly reduce the possibility of root

fracture and decementation (Morgano et al 1993).

The objective is to create a “cement-post-core” system with homogeneous properties

and physical characteristics similar to tooth tissues. To fulfill esthetic requirements,

quartz- and glass-fiber posts embedded in a filled resin matrix have been developed.

According to the manufacturer, the mechanical properties of these posts are similar to

those of carbon posts and provide an additional esthetic benefit. Fiber posts appear to

be biocompatible, are easy to insert, and are time and cost effective. Moreover, there

is no need for temporary fillings, since the post is placed using a one-stage technique.

The system is conservative with regard to the remaining dental structure and offers

the possibility of orthograde retreatment in cases of endodontic failure (de Rijk

2000). Several in vitro studies have been conducted on carbon- fiber posts, although

only a few investigations have been carried out on the esthetic quartz-fiber post

system. Relatively few reports have evaluated the in vivo effectiveness of carbon-

and quartz-fiber posts.

116

A 6-year clinical study using carbon-fiber posts (Composiposts) (Dallari et al 1996)

reported two failures in 575 restorations. A retrospective study (Fredrikson et al

1998) reported that the Composipost system performs favorably after 2 to 3 years. In

fact, only 2% of treated teeth had to be extracted, and none of these failures were

attributable to the fiber post system. Others (Ferrari et al 2000) evidenced a 3.2%

failure rate among 1,304 treated teeth restored with carbon- and quartzfiber posts in a

6-year study. Neither of these studies reported root fractures. The purpose of the

present prospective clinical follow-up was to evaluate the survival rate of 180

endodontically treated teeth restored using quartz-fiber posts and composite resin

material, and finalized with metal-ceramic or all-ceramic crowns over a 30- month

period.

Materials and Methods

One hundred eighty endodontically treated teeth in 132 patients (aged 18 to 65 years)

were restored by 13 different operators (Table 1). Restored teeth had the following

characteristics: need for prosthetic crown, root canal therapy performed at least 3

months previously with no subjective or objective symptoms, and no lesions visible

upon radiography. Radiographs were taken with the long-cone technique when the

restoration was performed and were examined with approximately 5_ magnification.

Teeth were restored with Æstheti-Plus quartz-fiber posts (RDT) (Figs 1 and 2). This

post system is composed of equally aligned quartz fibers that are longitudinally

embedded in an epoxy resin matrix. Posts are available in three sizes, have a

cylindric, double-section shape, and are 22 mm long. Post 1 has a diameter of 1.4

mm in the wider cylindric section and 1.0 mm at the narrow end, post 2 has a

diameter of 1.8 mm at the large end and 1.2 mm at the narrow portion, and post 3 has

diameters of 2.1 mm and 1.4 mm, respectively. This special shape is apparently

better adapted to the prepared canal. The choice of three different diameters provided

the possibility to find the adequate post dimension following the criteria of maximum

conservation of the residual dental tissue. In accordance with previous studies

(Fredrikson et al 1998 Chalifoux 1998) the following parameters were considered

relevant: number of canals; remaining tooth tissue, defined as complete (C; 66% to

100% of the tooth), partial (P; 33% to 65% of the tooth), or absent (A; 0% to 32% of

117

the tooth), shape of the canal space; and tooth antagonist. Among the 180 treated

teeth, 69% percent presented one root canal, and 31% percent had two or three

canals. In 14 maxillary premolars and 11 molars (four maxillary and seven

mandibular ones), the restorations were done with the anchorage of two posts per

tooth. Thus, among the 180 teeth restored, 205 canals were treated, and the same

number of posts was used. Seventy-nine percent of the canals treated showed a round

shape (R), 19% an oval one (O), and 2% a semicircular shape defined as C. Of the

opposing occluding teeth, 53% were natural teeth, 19% had metal-ceramic crowns,

16% were dentures, 11% occluded with metal-resin crowns, and 1% were not in

occlusion.

Clinical Procedures

Prior to cavity preparation, a rubber dam was placed, and provisional restorations

were removed with 80- µm diamond burs (205, Intensiv) in a medium-speed

handpiece under water cooling. Following direct clinical observation and

radiographic examination, the operator selected the most suitably sized fiber post.

Root preparation was done with a Pre-forma Drill and a Forma Drill in a slow-speed

handpiece with water spray. These calibrated burs provided a uniform preparation

and a thin and equally distributed layer of resin surrounding the post after its

cementation. The post was then reduced to the proper length using an 80-µm

diamond bur (206, Intensiv) in a highspeed handpiece with water spray; the bur was

kept perpendicular to the long axis of the post to avoid damaging the fiber structure

and its mechanical characteristics.

The length of the post was at least equal to the length of the clinical crown, always

respecting the apical gutta percha seal of 4 mm. The root canal was treated with 32%

phosphoric acid (Bisco) for 15 seconds, rinsed with deionized water, and gently dried

with air and a paper cone to verify that no traces of acid remained in the root

preparation.

Equally mixed primers A and B (Bisco) were applied with a Superfine Microbrush

(Microbrush) in the canal and on the post surface and then gently dried to permit the

evaporation of the acetone. Pre-Bond (Bisco) was applied inside the canal with a

paper cone. Next, the two components of the self-curing C&B Resin Cement (Bisco)

118

were mixed and applied at the edge of the root and on the post, which was

immediately placed into the prepared canal. The excess cement was trimmed and

given an adequate setting time (Fig 3). Core buildup was then performed using Core-

Flo (Bisco) or Bis-Core (Bisco) self-curing resin composite. Clinical recalls were

performed at 6, 12, 24, and 30 months. At the last recall, all teeth were crowned.

Three dentists evaluated the clinical performance of the restored teeth. The observers

were not blinded. Outcome was considered successful if the post and core were in

situ with no displacement or detachment of the post, no crown or prosthesis

decementation, and no post, core, or root fracture. Subjective symptoms reported by

the patients were considered potential signs of failure.

Results

After a period of 30 months, three failures were recorded; all took place during the

temporary phases during removal of the resin temporary restoration. The first failure,

recorded after 2 weeks in a maxillary left first premolar, was a cohesive fracture at

the edge of the composite resin of the core and did not involve the post structure. The

remaining dental structure was classified as P (33% to 65%), the canal shape was

oval, the post used had a diameter of 1.4 mm (post 1), the length of the post was 8

mm (slightly longer than the clinical crown), and the antagonist was a natural tooth.

After the fracture, the restoration was immediately replaced and the case was

finalized, and the post was still working successfully after 30 months. The second

and the third failures were adhesive fractures involving the cement-post-core

detaching from the dentinal walls of the root canal. The failures involved two canines

(one maxillary and one mandibular). Both teeth presented a rather low remaining

tooth structure (A; 0% to 32%) and a very large canal diameter, probably because of

excessive preparation during the endodontic therapy. Number 2 posts, which have a

diameter of 1.8 mm in the upper portion, had been used; in spite of this, a thick layer

of cement was around the posts. The lengths of the posts were 11 mm and 9 mm,

respectively. The 9-mm post was somewhat shorter than the clinical crown. Its length

had been imposed by the short root and the necessity to leave an adequate seal of the

apex. The antagonist was a natural tooth in one case and a ceramic crown in the

other. In both cases, the post and core were replaced immediately. The cases were

finalized and were still successful as of this writing. No crown or prosthesis

119

decementations or post, core, or root fractures were recorded. No significant

variations in terms of health of periodontal tissues were observed. No caries were

detected, and no subjective symptoms were reported. The three failures represented

1.7% of all treated teeth, giving a cumulative survival rate of 98.3% in a Kaplan-

Meier survival curve.

Discussion

The implementation of a new, homogeneous post and- core system is a good starting

point for the creation of a reliable substructure for prosthetic rehabilitation. In

combination with resin adhesive systems, resin cements, and composite resins, fiber

posts present biomechanical characteristics similar to those of the dentin. To date, the

favourable clinical outcomes observed encourage more widespread use. In

combination with the resin cement and composite resin restoration, the clinical

performance of the Æstheti-Plus fiber posts was good over a 30-month period. The

causes underlying the three precocious failures of the three restorations were

examined in detail. In the first case, a cohesive fracture, a bubble embedded in the

composite resin was found. The defect may have been due to insufficient

polymerization, possibly because of operator error. In the second and third cases,

adhesive fractures between the cement and the dental tissue were found. These

failures might also have been related to a procedural error by the operators or to the

excess layer of cement around the post, which may have provoked a change in the

mechanical behaviour of the cement-post–composite resin complex. The resin

cements that are supposed to provide the best adhesion (Ross et al 1991, Goldstein et

al 1986) may represent a weak point in the system, mostly when a very thick and non

uniform cement layer surrounds the post. It is reasonable to further detail the

mechanical procedures that the operator may perform during the temporary phases

and relate these to possible failures. It should be underlined that the cases showing

the adhesive failures did not have the 2-mm ferrule of dentin considered to be

important to obtain a high success rate (Assif et al 1991). In accordance with other

clinical studies, the observed success rate of 98.3% was high and no root fractures

120

occurred. Therefore, it was possible to replace the failing restorations (Ferrari et al

2000). Other desirable properties of these posts are their biocompatibility and

resistance to corrosion (Torbjörner et al 1996, Jockisch et al 1992). Fredrikson

(Fredrikson et al 1998) reported no differences in the periodontal conditions between

carbon fiber–treated teeth and controls. Radiographic examination of bone height

measured from the apex to the bone margin mesially and distally showed differences

on the mesial side, but not on the distal surface, between the treated and control teeth.

Other in vitro studies have compared fiber posts with traditional metal posts to

determine which system offers the best mechanical properties. Purton (Purton et al

1996) compared fiber posts and stainless steel root posts by three-point bending tests

to derive the transverse modulus of elasticity of the posts. The carbon-fiber material

was stiffer under transverse loading than stainless steel and thus appears to have

adequate rigidity for its designed purpose. This higher rigidity allows smaller

diameters to be used with equivalent strength, in accordance with the conservative

principles. One weak point of the system is the bond between the post and the

composite core material. The authors therefore suggested that a surface treatment or

modification in configuration should be introduced. Drummod (Drummod 2000)

evaluated the pullout (shear) strength of stainless steel posts and three different fiber

posts and found no significant differences. With respect to flexure strength, all fiber

post systems showed a significant decrease following thermocycling, probably

because of the degradation of the polymer holding the fibers together and/or to the

fibers themselves. The clinical relevance of this data still needs to be investigated.

Relatively few clinical studies examining the success and failure of metallic posts

have been reported. Sorensen (Sorensen et at 1984) reported an 8.6% failure rate

resulting from post dislodgments, root fractures, or post perforations. Others reported

a 6.5% failure rate after 10 years or more (Weine et al 1991) and an 8.3% frequency

of failure after 2 to 3 years (Torbjörner et al 1995). Lewis (Lewis et al 1988) stated

that failures of the post and core are more likely to occur within the first 3 years of

cementation. A retrospective study (Ferrari et al 2000) evaluated the outcome of cast

posts and cores and carbon-fiber posts after 4 years of clinical service. Ninety-five

percent of the teeth restored with fiber posts showed clinical success, 3% were

excluded, and 2% showed endodontic failure. Among the teeth restored with cast

posts, 84% showed clinical success, 9% showed root fracture, 3% showed

121

endodontic failure, 2% had dislodgment of the post or crown, and 2% were excluded.

Statistical evaluation has thus indicated that carbon-fiber posts are superior to

conventional cast posts. In accordance with this study, the failures recorded with

fiber posts were more benign. When a vertical fracture occurs, the entire element

must be extracted. When finalized with all-ceramic crowns, the strong demand for a

post-and-core complex with good esthetic results has guided both researchers and

manufacturers toward the introduction of restoring systems that meet these

requirements. According to a previous study, the clinical behavior of carbon-fiber

posts and quartz-fiber posts is equivalent (Chalifoux 1998). To determine how much

the post aspect will influence light transmission under the thickness of the ceramic,

the following guidelines can be used (Vichi et al 2000). With a 2-mm ceramic

thickness, color differences are clinically irrelevant, while with a ceramic thickness

of 1.5 mm, the color differences are clinically visible in some cases and detectable

only with a spectrophotometer in others. With a ceramic thickness of less than 1.5

mm, the aspect of the post-and-core restoration influences the all-ceramic crown in a

manner that may be clinically unacceptable. The fiber post system offers a time

savings, simple clinical procedures, and reliable results. Moreover, the technique is

less invasive than other post systems because of the shorter length required, and thus

the apical seal can be left around 4 mm (Goerig et al 1983a, Goerig et al 1983b). In

cases of endodontic failures, it is possible to remove the post with the specific bur

(Sakkal 1996). It is obvious that the clinician must follow the manufacturer’s

instructions while treating the remaining dental tissues. The use of a rubber dam is

imperative while performing the restoration. The combination of good mechanical

performance with satisfactory esthetics may be a good starting point toward the

improvement of routine dentistry and toward predicting success. We are awaiting

future evaluations of these encouraging data.

Within a 30-month period, 205 Æstheti-Plus quartz-fiber posts were used to restore

180 teeth with clinical success. The three failures recorded represented 1.7% of

treated teeth. All three failures occurred during the temporary phases. The cohesive

fracture and two adhesive fractures involved only the post-and-core restoration. No

root or post fractures occurred. In the case of failure, it was possible to replace the

restoration without losing the element.

122

Table 1 Distribution of Treated Teeth According to Type Central Lateral Jaw

Central incisor

Lateral incisor

Canines

Premolars

Molars

Maxilla

43

17

28

24

8

Mandibole

9

11

16

8

16

Both

52

28

44

32

24

Fig 1 Æstheti-Plus quartz-fiber post cemented in a maxillary

right central incisor.

Fig 2 Completed core buildup.

References

Assif D, Pilo R, Marshak B. Restoring teeth following crown lengthening procedures. J Prosthet Dent 1991;65:62–64.

Chalifoux PR. Esthetic restoration of endodontically treated teeth: Factors that

affect prognosis. J Esthet Dent 1998;10:75–83.

Dallari A, Rovatti L. Six years of in vitro/in vivo experience with Composipost. Compendium 1996;17:57–63.

de Rijk WG. Removal of fiber posts from endodontically treated teeth. Am J

Dent 2000;13:19b–21b.

Drummond JL. In vitro evaluation of endodontic posts. Am J Dent 2000;13:5b–8b.

Duret B, Reynaud M, Duret F. Un nouveau concept de reconstituction corono-radiculaire: Le Composipost (1). Chir Dent France 1990;540:131–141.

123

Duret B, Reynaud M, Duret F. Un nouveau concept de reconstituction corono-radiculaire: Le Composipost (2). Chir Dent France 1990;542:69–77.

Ferrari M, Vichi A, García-Godoy F. Clinical evaluation of fiberreinforced epoxy resin posts and cast post and cores. Am J Dent 2000;13:15b–18b.

Ferrari M, Vichi A, Mannocci F, Mason PN. Retrospective study of the clinical performance of fiber posts. Am J Dent 2000;13:9b–13b.

Fredrikson M, Astback J, Pameius M, Arvidson K. A retrospective study of 236 patients with teeth restored by carbon fiber reinforced epoxy resin posts. J Prosthet Dent 1998;80:151–157.

Goerig AC, Mueninghoff LA. Management of the endodontically treated tooth.

Part II: Technique. J Prosthet Dent 1983;49:491–497.

Goerig AC, Mueninghoff LA. Management of the endodontically treated tooth.

Part I: Concept for restorative designs. J Prosthet Dent 1983;49:340–345.

Goldstein GR, Hudis SI, Weintraub DE. Comparison of four techniques for the cementation of posts. J Prosthet Dent 1986;55: 209–211.

Goodacre CJ, Spolnik KJ. The prosthodontic management of endodontically

treated teeth: A literature review. Part 1: Success and failure data, treatment concepts.

J Prosthodont 1994;3:243–250.

Jockisch KA, Brown SA, Bauer TW, Merritt K. Biological response to chopped-carbon-fiber-reinforced peek. J Biomed Mater Res 1992;26:133–146.

Lewis R, Smith BG. A clinical survey of failed post retained crowns. Br Dent J 1988;165:95–97.

Morgano SM, Milot P. Clinical success of cast metal posts and cores. J Prosthet Dent 1993;70:11–16.

Purton DG, Payne JA. Comparison of carbon fiber and stainless steel root canal

posts. Quintessence Int 1996;27:93–97.

Ross RS, Nicholls JI, Harrington GW. A comparison of strains generated during placement of five endodontic posts. J Endod 1991; 17:450–456.

Sakkal S. Carbon-fiber post removal technique. Compend Contin Educ Dent Suppl 1996;(20):86S.

Sorensen JA, Martinoff JF. Clinically significant factors in dowel design. J Prosthet Dent 1984;52:28–35.

Torbjörner A, Karlsson S, Ödman PA. Survival rate and failure characteristics for two post designs. J Prosthet Dent 1995;73:439–444.

Torbjörner A, Karlsson S, Syverud M, Hensten-Pettersen A. Carbon fiber reinforced root canal posts. Mechanical and cytotoxic properties. Eur J Oral Sci 1996;104:605–611.

Trabert KC, Cooney JP. The endodontically treated tooth. Restorative concepts and techniques. Dent Clin North Am 1984;28: 293–951.

Vichi A, Ferrari M, Davidson CL. Influence of ceramic and cement thickness on

the masking of various type of opaque posts. J Prosthet Dent 2000;83:412–417.

Weine FS, Wax AH, Wenckus CS. Retrospective study of tapered, smooth post system in place for 10 years or more. J Endod 1991;17: 293–297.

124

125

Chapter 11 summary and conclusions

Until now, clinical application of dental restorations to restore missing posterior teeth

could only be performed using conventional metal systems because of high

mechanical loading in this area. The resin-bonded fixed partial denture (FPD) is a

treatment alternative for replacement of missing teeth when conservation of tooth

structure is needed (Freilich et al 1998). Currently, alternative systems like fibre-

reinforced composite (FRC) dental materials, are recommended for their ability to

withstand these mastication forces. Fibre-reinforced materials combine the basically

different mechanical properties of fibres and a matrix, in which the fibres are

embedded. The fibres demonstrate high tensile strength, a high tensile modulus, and

low shear strength, while the matrix is characterized by high toughness. In an

optimum fibre-reinforced material, the tensile strength of the fibre is combined with

the high toughness of the matrix (Vallittu et al 1999a, Vallittu et al 1999b). Fibre-

reinforced technology is used wherever high stress occurs and low weight is required,

such as in the aeronautical and shipbuilding industries. Numerous attempts have been

made to develop ceramic systems that eliminate metal infrastructures and provide

optimal distribution of reflected light. Currently clinicians have an increasing range

of ceramics capable of delivering high quality aesthetic restorations to choose from

for many clinical indications. New ceramic systems involve reinforced ceramic cores

through dispersion with leucite (Kon et al 1994, Seghi et al 1995, Denry et al 1996,

Mackert et al 1996, Mackert et al 2000), glass infiltration into sintered alumina

(Al2O3) (Probster et al 1994, Sadoun et al 1994), the use of high-purity alumina

(Andersson et al 1993) or zirconium dioxide (zirconia, ZrO2) (Piconi et al 1999).

Conclusions and recommendations

The following conclusions and recommendations may be drawn from our basic and

clinical evaluations on the use of FRC and all ceramic systems applied in partial

restorations:

126

1) Excluding the therapy with dental implants, inlay fixed partial dentures

represent the most conservative treatment for the replacing of the missing

tooth

2) The design of the FRC framework can play important role in supporting the

layering material in a continuous manner. Vestibular and buccal FRC

extensions in the pontic element can increase the bonding area between the

framework and resin composite as well as hold the veneer material better

during occlusal loading.

3) The flexibility of the framework may play an important role in the marginal

adaptation of IFPDs and more rigid materials could transfer the stress to the

margin to a smaller degree than flexible materials

4) The different arrangement of the glass fibers can increase the fracture

strength of FRC three-unit bridges and IFPDs. Increased fracture strength

could be useful in case of long span framework or short clinical abutments

5) fiber reinforced composites exhibit good overall mechanical properties and

the related failure modes are usually more conservative and favorable when

compared to the traditional metal-ceramic three unit bridges.

6) In short observation period the IFPDs in combination with a 3-step adhesive

demonstrated good clinical service but we need more observations in order

to create a sound basis for the final assessment of IFPDs

7) Longer observation periods and comparisons with new reinforced ceramic

materials will help to identify the most suitable choice of material for IFPDs

References

Kon M, Kawano F, Asaoka K, Matsumoto N. Effect of leucite crystals on the strength of glassy porcelain. Dent Mater 1994;13:138–47. Seghi RR, Sorensen JA. Relative flexural strength of six new ceramic materials. Int J Prosthodont 1995;8:239–46. Seghi RR, Denry IL, Rosenstiel SF. Relative fracture toughness and hardness of new dental ceramics. J Prosthet Dent 1995;74:45–50. Denry IL, Mackert Jr. JR, Holloway JA, Rosenstiel SF. Effect of cubic leucite stabilization on the flexural strength of feldspathic dental porcelain. J Dent Res 1996;75:1928–35. Mackert JR, Russell CM. Leucite crystallization during processing of a heat-pressed dental ceramic. Int J Prosthodont 1996;9:261–5.

127

Mackert Jr JR, Williams AL, Ergle JW, Russell CM. Water-enhanced crystallization of leucite in dental porcelain. Dent Mater 2000;16:426–31. Probster L, Diehl J. Slip-casting alumina ceramics for crown and bridge restorations. Quintessence Int 1992;23:25–31. Sadoun M, Asmussen E. Bonding of resin cements to an aluminous ceramic: a new surface treatment. Dent Mater 1994;10:185–9. Andersson M, Oden A. A new all-ceramic crown—a dense-sintered, high purity alumina coping with porcelain. Acta Odontol Scand 1993;51:59–64. Piconi C, Maccauro G. Zirconia as a ceramic biomaterial. Biomaterials 1999;20:1–25. Sadoun M, Asmussen E. Bonding of resin cements to an aluminous ceramic: a new surface treatment. Dent Mater 1994;10:185–9.

Freilich MA, Duncan JP, Meiers JC, Goldberg AJ. Preimpregnated, fiber-reinforced prostheses. Part I: basic rationale and complete coverage and intracoronal fixed partial denture designes. Quintessence Int 1998;29:689-696.

Vallittu PK. Flexural properties of acrylic resin polymers reinforced with

unidirectional and woven glass fibers. J Prosthet Dent 1999a;25:100-105. Vallittu PK. Prosthodontic treatment with a glass fiber-reinforced resin-bonded

fixed partial denture: a clinical report. J Prosthet Dent 1999b;82:132-135.

List of papers part of this thesis Monaco C, Ferrari M, Miceli GP, Scotti R. Clinical Evaluation of Fiber-Reinforced

Composite IFPDs. 2003 Int J Prosthodont 16:319-325.

Monaco C, Miceli GP, Scotti R. Die mit dem neuen, mikrogefüllten Komposit-

Material SR Adoro verblendete Inlay-Brücke. Quintessenz Zahntech 29,3,292-305

(2003).

Malferrari S, Monaco C, Scotti R. Clinical evaluation of teeth restored with quartz

fiber-reinforced epoxy resin posts. 2003 Int J Prosth 16;39-44.

Monaco C, Ferrari M, Caldari M, Baldissara P, Scotti R.

Randomized controlled trial of Fiber-Reinforced Composite Inlay Fixed Partial

Dentures: two-year results. Int J Prosth 2006, n press.

Monaco C, Krejci I, Bortolotto T, Perakis N, Ferrari M, Scotti R.

128

Marginal adaptation of one Fiber Reinforced Composite and two different all-

Ceramic Inlay Fixed Partial Denture Systems. Int J Prosth 2006, in press.

Marginal adaptation of one Fiber Reinforced Composite and two different all-

Ceramic Inlay Fixed Partial Denture Systems. Int J Prosth 2006, in press.

Baldissara P, Valandro LF, di Fiore P, Monaco C, Ferrari M, Bottino MA, Scotti R.

Fatigue test in shear: its effect on bond of a glass-infiltrated alumina ceramic to

human dentin, using different luting procedures. J Adhes Dent 2006, in press.

List of abstract part of this thesis Monaco C, Ferrari M . Randomized clinical trial of inlay fixed partial dentures .

Abstr 585 119CED-NOF of the IADR meeting Amsterdam 14-17 september 2005.

Monaco C. Adhesive inlay bridges with FRC: clinical and laboratory investigations.

Abstr SO460C 119CED-NOF of the IADR meeting Amsterdam 14-17 september

2005.

C. Monaco, P. di Fiore, P. Baldissara, R. Scotti. Fatigue resistance of in-

Ceram/Dentin bonding: a comparative study. 2003 J Dent Res 82. Special issue B

Abstr 1462.

Monaco C, Baldissara P, Caldari M, Scotti R. Six-month evaluation of post

operative sensitivity in IFPDs. J Dent Res Abstr 579. Honolulu 2004

Baldissara P., Melilli D, Monaco C, Ciocca L. Fatigue resistance of fiber posts and

different build-up materials. J Dent Res Abstr 579. Honolulu 2004.

Baldissara P, Quadrelli M, Monaco C, Scotti R. E-Modulus and Flexure Strength of

FRC: Introducing Boron Fibers. Abstr 372 119CED-NOF of the IADR meeting

Amsterdam 14-17 september 2005.

129