Sunteți pe pagina 1din 67

Ofcial Publication of the American Academy of Esthetic Dentistry, Scandinavian Academy of Esthetic Dentistry, Japan Academy of Esthetic Dentistry,

International Federation of Esthetic Dentistry, American Academy of Cosmetic and Adhesive Dentistry, Australian Society of Aesthetic Dentistry and the Belgian Academy of Esthetic Dentistry

2000

Volume 12

Number 6

Journal of Esthetic Dentistry


ADMINISTRATIVE BOARD EDITOR-IN-CHIEF

Harald O. Heymann, DDS, MEd


ASSOCIATE EDITOR

Edward J. Swift Jr., DMD, MS


EDITORIAL ASSISTANT

Betty T. Cates
SECTION EDITORS

Edward P. Allen, DDS, PhD Stephen C. Bayne, MS, PhD Terence Donovan, DDS, MS Vincent G. Kokich, DDS, MSD Jorge Perdigo, DMD, MS, PhD Louis F. Rose, DDS, MD Edwin S. Rosenberg, BDS, H.Dip.Dent, DMD Henri Rotsaert, RDT Richard E. Walton, DDS, MS
EDITORIAL ADVISORY BOARD

Leonard Abrams (USA) Steve Acquilino (USA) Pinhas Adar (USA) Ken Anusavice (USA) Donald Arens (USA) Luiz Narciso Baratieri (Brazil)

William Becker (USA) Urs Belser (Switzerland) Gunnar Bergenholtz (Sweden) Robert Berger (USA) Joe Camp (USA) Gerard J. Chiche (USA) Noah Chivian (USA) Gordon J. Christensen (USA) E. Steven Duke (USA) Jim Dunn (USA) Jack Ferracane (USA) Mauro Fradeani (Italy) Mark J. Friedman (USA) David A. Garber (USA) Jaime A. Gil (Spain) Ronald E. Goldstein (USA) Ueli Grunder (Switzerland) Van B. Haywood (USA) Abraham Ingber (USA) John Kanca (USA) Ivo Krejci (Switzerland) Masahiro Kuwata (Japan) Paul Lambrechts (Belgium) Karl F. Leinfelder (USA) William Liebenberg (Canada) Felix Lutz (Switzerland) Pascal Magne (Switzerland) Ronald I. Maitland (USA)

Lloyd Miller (USA) Preston D. Miller (USA) Ivar A. Mjr (Norway/USA) Dan Nathanson (USA) Linda C. Niessen (USA) Patrick Palacci (France) Nicola Pietrobon (Switzerland) William R. Proffit (USA) Franois Roulet (Germany) Cliff Ruddle (USA) Frederick A. Rueggeberg (USA) Fortunato Santos (Brazil) Peter Schrer (Switzerland) Cherilyn G. Sheets (USA) Richard Simonsen (USA) Dan Sneed (USA) Frank Spear (USA) Howard F. Strassler (USA) Jrg Strub (Germany) Daniel Sullivan (USA) Asami Tanaka (USA) Dennis P. Tarnow (USA) Martin Trope (USA) Arnold S. Weisgold (USA) John West (USA) Ray Williams (USA) David C. Winkler (Denmark) Robert Winter (USA)

OFFICIAL PUBLICATION OF THE AMERICAN ACADEMY OF ESTHETIC DENTISTRY JAPAN ACADEMY OF ESTHETIC DENTISTRY SCANDINAVIAN ACADEMY OF ESTHETIC DENTISTRY INTERNATIONAL FEDERATION OF ESTHETIC DENTISTRY AMERICAN ACADEMY OF COSMETIC AND ADHESIVE DENTISTRY AUSTRALIAN SOCIETY OF AESTHETIC DENTISTRY BELGIAN ACADEMY OF ESTHETIC DENTISTRY

The JOURNAL OF ESTHETIC DENTISTRY (ISSN 1040-1466) is published bimonthly in January, March, May, July, September, and November by BC Decker Inc, 20 Hughson St. South, 10th Floor, P.O. Box 620, L.C.D. 1, Hamilton, Ontario L8N 3K7. The annual subscription rate for the U.S. is $175.00 U.S. for individuals and $269.00 U.S. for libraries and institutions. The foreign rate is $225.00 U.S. for individuals and $299.00 U.S. for libraries and institutions. Single issues are available in the U.S. and Canada for $52.00 U.S.; foreign, $58.00 U.S. Claims for missing issues can be honored only up to 3 months for domestic addresses, 6 months for foreign addresses. Duplicate copies will not be sent to replace ones undelivered through failure to notify BC Decker Inc of change of address. Advertising inquiries should be addressed to John Birkby, BC Decker Inc, 20 Hughson St. South, P.O. Box 620, L.C.D. 1, Hamilton, Ontario, Canada L8N 3K7. Tel: (905) 522-7017; Fax: (905) 522-7839; in Canada and U.S.: 1-800-568-7281. Address subscription inquiries to BC Decker Inc, P.O. Box 620, L.C.D. 1, Hamilton, Ontario L8N 3K7, Tel: (905) 522-7017. Customer Service Tel: 1-800-568-7281. No responsibility is assumed by the Publisher for any injury and/or damage to persons or property as a matter of product liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. No suggested test or procedure should be carried out unless, in the readers judgment, its risk is justified. Because of rapid advances in the medical sciences, we recommend that the independent verification of diagnoses and drug dosages should be made. Discussions, views, and recommendations for medical procedures, choice of drugs, and drug dosages are the responsibility of the authors. The appearance of advertising in the JOURNAL OF ESTHETIC DENTISTRY does not constitute a guarantee or endorsement of the quality or value of such product or of the claims made for it by its manufacturer. The fact that a product, service, or company is advertised in the JOURNAL OF ESTHETIC DENTISTRY shall not be referred to by the manufacturer in collateral advertising. 2000 BC Decker Inc. All rights reserved.

287

PERSPECTIVES

Title???
Fred Rueggeberg, Murray Bouschlicher CLINICAL ARTICLES ?? RESEARCH ARTICLES ??

291

Developments in Shrinkage Control of Adhesive Restoratives


Bibi S. Dauvillier, MSc, Maxim P. Aarnts, PhD, Albert J. Feilzer, DDS, PhD

300

Curing Dental Resins and Composites by Photopolymerization


Jeffrey W. Stansbury, PhD

309

Polymerization Contraction Stress of Resin Composite Restorations in a Model Class I Cavity Configuration Using Photoelastic Analysis
Yoshifumi Kinomoto, DDS, PhD, Mitsuo Torii, DDS, PhD, Fumio Takeshige, DDS, PhD, Shigeyuki Ebisu, DDS, PhD

320

Effect of Distance on the Power Density from Two Light Guides


Richard B. Price, DDS, MS, FDS RCS (Edin), Tore Drand, DDS, PhD, Mary Sedarous, BSc, Pantelis Andreou, PhD, Robert W. Loney, DMD, MS

328

Effect of Ramped Light Intensity on Polymerization Force and Conversion in a Photoactivated Composite
Murray R. Bouschlicher, DDS, MS, Frederick A. Rueggeberg, DDS, MS

340

Polymerization Depth of Contemporary Light-Curing Units Using Microhardness


Frederick A. Rueggeberg, DDS, MS, Janet W. Ergle, Donald J. Mettenburg

350

INDEX

Journal of Esthetic Dentistry

BC DECKER INC

Business Information for Readers


Journal of Esthetic Dentistry
Owned and published by BC Decker Inc 20 Hughson Street South P.O. Box 620, LCD 1 Hamilton, ON L8N 3K7 Tel: 905-522-7017; Fax: 905-522-7839 E-mail: info@bcdecker.com Website: http://www.bcdecker.com Editor-in-Chief Harald O. Heymann, DDS, MEd Chapel Hill, North Carolina Tel: 919-966-2770; Fax: 919-966-5660 E-mail: harald_heymann@dentistry.unc.edu Editorial Assistant Betty T. Cates Chapel Hill, North Carolina Tel: 919-843-9744; Fax: 919-966-5660 E-mail: betty_cates@dentistry.unc.edu SUBSCRIPTION RATES The 2000 annual subscription rates are: United States and Canada (U.S.$): individual, $175.00; institution, $269.00; single issue, $52.00. Elsewhere (U.S.$): individual, $225.00; institution, $299.00; single issue, $58.00. For subscriptions outside the USA or Canada, payment must be in U.S. funds drawn on a U.S. bank free of bank charges, or a U.S. dollar International Money Order free of service fees. VISA, MasterCard, and American Express are also accepted. Airmail rates for delivery overseas are available on request. Remit payments and correspondence to: BC Decker Inc, Customer Service Department, P.O. Box 620, LCD 1, Hamilton, ON L8N 3K7 Canada, or P.O. Box 785, Lewiston, NY 14092-0785 USA. Tel: 1-800-568-7281 (USA and Canada), 905-522-7017 (elsewhere), Fax: 905-522-7839, E-mail: info@bcdecker.com. Please include a mailing label from a recent issue for renewal orders. In Japan, contact Ishiyaku Publishers, Inc., 7-10 Honkomagome, I-Chome, Bunkyo-Ku, Tokyo 113, Japan. Tel: 03-5395-7631; Fax: 03-5395-7633; Telex: 2723298 MDP. CLAIMS Claims for issues not received or for damaged issues must be made within three months of publication for subscriptions mailing within North America, six months for destinations elsewhere. Issues claimed beyond these limits must be prepaid at the single issue rates listed below. Duplicate copies will not be supplied to replace those undelivered through failure to notify the Publisher of a change of address. Please let us know of address changes as soon as possible. Send us your current mailing label (with the old address), your new address, and the date the change becomes effective. BACK ISSUES Back issues are available at the single issue price of U.S. $52.00 in the USA and Canada, U.S. $58.00 elsewhere, or the complete volume at current volume prices. All prices are in U.S. dollars. Orders must be prepaid by one of the methods listed in the subscription information. Please check with Customer Service for availability at 1-800-568-7281 (USA and Canada), 905522-7017 (elsewhere), 905-522-7839 (Fax), or E-mail: info@bcdecker.com. BULK REPRINTS Reprints of Journal articles are available in minimum quantities of 100. To order, please contact the Sales Department, BC Decker Inc, at the address or phone number listed above. INDEXING The Journal is included in MEDLINE on the MEDLARS. DISPLAY ADVERTISING The appearance of advertisements in the Journal is not a warranty, endorsement, or approval of the products or services advertised, or of their effectiveness, quality, or safety. The Publisher, the American Academy of Esthetic Dentistry, the Japan Academy of Esthetic Dentistry, Scandinavian Academy of Esthetic Dentistry, International Federation of Esthetic Dentistry, American Academy of Cosmetic and Adhesive Dentistry, Australian Society of Aesthetic Dentistry, and the Belgian Academy of Esthetic Dentistry disclaim any responsibility for any injury to persons or property resulting from any ideas or products referred to in the articles or advertisements. CLASSIFIED ADVERTISING Rates are $735.00 for a 1/8 page ad, $945.00 for a 1/4 page ad, $1,155.00 for a 1/3 page ad, $1,610.00 for a 1/2 page ad, and $2,310.00 for a 1 page ad. For further information, contact the Sales Department at BC Decker Inc at the address or phone number listed above. DISCLAIMER The statements and opinions in the articles of this Journal are solely those of the individual authors and contributors and not of the Publisher or the Societies that list the Journal as their official publication. COPYRIGHT Material printed in the Journal is covered by copyright. All rights reserved. Except under circumstances within fair use as defined by copyright law, no part of this publication may be reproduced, displayed, or transmitted in any form or by any means, electronic or mechanical, including photocopying or by any information storage and retrieval system, without the prior written permission of the copyright owner, BC Decker Inc. Requests should be made to Paula Presutti, Rights and Permissions, BC Decker Inc, 20 Hughson Street South, P.O. Box 620, LCD 1, Hamilton, ON L8N 3K7; Fax: 905-522-7839. MICROFORM The Journal is available on Microfilm from Bell and Howell Information and Learning, 300 N. Zeeb Road, Ann Arbor, MI 48106; 415-433-5500.

PUBLISHER BC Decker Inc President, Brian C. Decker Vice President, Finance, Doug M. Fleming Director, Sales and Marketing, David LeGallais Production Editor, Paula Presutti Rights and Permissions, Paula Presutti Customer Service Manager, Cathy Calvin Distribution and Postal Affairs, Deborah J. Maraldo

ADVERTISING John Birkby BC Decker Inc 20 Hughson Street South P.O. Box 620, LCD 1 Hamilton, ON L8N 3K7 Tel: 905-628-4309; Fax: 905-628-6847 E-mail: jbirkby@cableworks.org

Perspectives

TITLE? Light-activation of dental restorative materials has become a way of life for the average clinician. It provides the bread and butter of many practices, and enables immediate, direct, and highly esthetic results to be obtained with minimal loss of patient time from work or school, and minimal chairside-contact time. In short, it has been a revolution many contemporary clinicians have lived through. Younger practitioners, however, may not fully appreciate the journey from chemical-cured composites to the light-cured generation. Gone are the days of hurried placement and crude, rushed shaping within a confined, distorted matrix band. Available now are materials offering setting-on-command, with prolonged placement time to sculpt, adapt, and contour the restoration and then, almost instantly, freeze it in place. Light-activation has improved esthetics (less restoration yellowing) and decreased porosity compared that which was inevitable from mixing of pastepaste formulations. However, we now know that these internal porosities actually aided to decrease stress development, and the slower rate of curing permitted the partially cured restoration to flow from the unbonded surfaces, relieving stress development at the toothrestoration interface. Thus, experience has come full circle. With proven clinical longevity to support its use, light-curing of dental restorations has become a predictable, routine practice. However, the revolution apparently is not over yet. Recent developments in curing light technology have shaken the complacence of many practitioners, and brought to light aspects of the polymerization process here-to-fore either ignored or not realized. Work performed years ago on the possible clinical implications of the rates at which resins cure and the potential for stress development were read but not truly appreciated for their insight and implication. With the advent of high speed curing and soft-start polymerization, the clinician is faced with the dilemma of jumping on the bandwagon to minimize precious chairside time, while attempting to provide restorations of high quality and long clinical durability. The breadth of choice for a light-curing unit is now bewildering. Claims are made from marketing aspects, with assumed or unfounded clinical data to support them. One needs only to scan product advertisement periodicals to comprehend the scope of curing units available and methods being advocated for polymerization: conventional cure, quartz tungsten halogen (QTH) sources, highintensity QTH, ramped output, stepped output, pulse-delay methods, argon ion lasers, plasma arc units (PAC lights), and the new light emitting diodes (LEDs). To assess proponent claims of the various methods of composite photopolymerization, the fundamentals underlying polymerization need to be understood. This understanding must be coupled with laboratory and clinical evidence supporting principals established from basic science. This special issue on photocuring is not meant to provide the final answer to these questions but instead is meant to be a framework upon which readers can start to

VOLUME 12, NUMBER 6, 2000

287

PERSPECTIVES

comprehend the complexities underlying what can so easily be taken for granted: shine the light, and it will cure. The articles contained in this issue represent a compilation of historic perspectives, basic science, and clinical realities of only a small sampling of the totality of issues related to this topic. As co-editors, we take pride in the assemblage of this work, because it represents the input of many talented and dedicated individuals who are experts in their own right. This issue has been developed with the clinician in mind. To enhance the profession, the clinician must become aware of issues. Awareness itself does not lead to change; it is only through understanding that change can arise. Thus, each author was specifically asked to present his or her work with the clinician in mind and, whenever possible, to emphasize the clinical relevance of the findings. Again, this issue is not THE answer, but is designed to take a few steps toward educating the practitioner so that well-informed, knowledgeable decisions can be made with respect to use of photocuring in an individuals practice. We are indebted to Dr. Harald Heymann for the encouragement, support, and motivation to provide this accumulation of works for the readers consideration. Fred Rueggeberg and Murray Bouschlicher Guest Editors

A RT I C L E S I N T H I S I S S U E Dauvillier, Aarnts, and Feilzer review the underlying causes of contraction of polymeric restorative materials, and the various factors that influence them. Some factors affecting stress development are out of the clinicians control (e.g., composite monomeric and filler formulation); however, the methods used for placement and light-curing are directly under the clinicians control. This article stresses the importance of knowing the relation of these manipulative factors to the development of stresses from polymerization. The problems of balancing a low polymerization reaction rate with obtaining a high, final conversion value to provide clinically adequate properties to the restoration are addressed. Allowing a composite to flow prior to reaching high moduli values (the gel point) permits a relief of these stresses instead of allowing them to build up within the material and at the materialtooth interface. Building up a composite in increments helps to reduce the configuration factor (C-factor) and minimizes stress development. The true benefits and mechanisms of stress alteration from placing low modulus resin liners is still not definitive, and the effects of variation in light intensity (the soft-start techniques and the new high output intensity models) still have not been proved. The influence of water sorption on stress relaxation of composites is also addressed. Clinicians may take the mechanisms underlying photocuring of light-activated dental restorations for granted. However, this process is complex, and knowledge of the components, the mechanisms underlying the reaction, and factors affecting the efficiency of the

288

JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG AND BOUSCHLICHER

potential for cure all have clinical significance. The fundamentals of this reaction as well as many other aspects of this topic are addressed by Stansbury in his work entitled Curing Dental Resins and Composites by Photopolymerization. The reasons underlying the mechanisms of oxygen inhibition, the difficulty in adequately curing some composite systems containing the new alternative photoinitiator systems, the influence of fillers in composites, the level of heat generated during composite cure, and the rationale underlying post-cure heating of composites are clearly addressed. Using a simulated Class I situation, Kinomoto and colleagues have developed a sophisticated but visual method of comprehending the magnitude and direction of stresses developed in a light-cured composite material, as well as in the material to which it is bonded. Interestingly, when the light-cured material was exposed in preparations made in extracted bovine teeth, gap formation was present. When this gap formed, low stresses showed in the cured composite, because the composite was not bonded to an interface. The distribution of stress patterns in model composite preparations was similar for both the self-cured and light-cured composites. However, the light-cured material demonstrated significantly higher stress values. Stresses also develop in the substrate that is being restored, as is clearly demonstrated in this work. This article demonstrates the location and variation of stress development along the lateral restoration walls as well as across the pulpal floor of a restoration. Clinicians can easily understand how and where stresses develop in resinbased restorations, and can then refine their techniques to minimize the deleterious effects of these forces. The influence of light-curing tip design on the power density delivered to the restorative material surface affects the overall potential for curing of the restoration, and its clinical durability. Thus, Price and colleagues, evaluated the effect of tip distance and power density between a standard 8-mm diameter curved light guide and an intensifying tip (the Turbo tip) in Effect of Distance on the Power Density from Two Light Guides. Even though the absolute intensity delivered when using the Turbo tip is greater than with the conventional tip at the distal tip end, the rate at which the power density declines as the tip distance increases occurred at a significantly higher rate with the Turbo tip than the conventional guide. Interestingly, at tip-to-target distances greater than 5 mm, the conventional light guide provided significantly greater power density than did the Turbo tip. Thus, clinicians need to be familiar with any focusing effect of light guides to use them to their greatest clinical potential. The rate at which composite polymerizes to bonded substrates may significantly affect the development of stresses between these materials, and thus affect the potential for successful bonding. Bouschlicher and Rueggeberg address these issues in their article Effect of Ramped Light Intensity on Polymerization Force and Conversion in a Photoactivated Composite. Using bonded surfaces attached to a force measuring device, the real-time force development between a commercially available composite and the bonded substrates was determined when using a light-curing unit in a variety of exposure modes: standard,

VOLUME 12, NUMBER 6, 2000

289

PERSPECTIVES

continuous output; an exponential increase in light intensity; a two-step method; and at half intensity of the continuous exposure. The results indicate that use of the exponential ramp significantly reduces stress development during curing and also results in equivalent total cure values to those achieved when the composite was exposed using standard techniques. Thus, laboratory data indicate that use of this type of soft-start cure method does verify theoretic concepts of stress reduction when a low rate of light intensity is used during the initial phases of exposure. The potential for a wide variety of commercially available light-curing units and techniques to polymerize a photoactivated composite was evaluated by Rueggeberg, Ergle, and Mettenburg in Polymerization Depths of Contemporary Light-Curing Units Using Microhardness. The same shade and lot of composite was exposed to conventional quartz tungsten halogen units, high-intensity QTH lights, soft-start techniques (step and pulse-delay modes), plasma arc (PAC) lights, and various exposure durations of an argon ion laser. The claims of curing light manufacturers and advocates of certain curing methods were evaluated on their ability to provide hardness of composite similar to that obtained with the standard technique: a 40-second exposure from a conventional QTH light source. Clinical usefulness of the results focuses on evaluating the ability of newer light-curing units to shorten exposure duration or to provide enhanced depth of cure compared with the conventional curing methods. In most instances, cure depths, using manufacturers recommended exposure durations, did not exceed the 2-mm depth traditionally accepted. Some high-intensity units (high-intensity QTH, the PAC light, and the laser) did provide for lowered exposure duration while presenting similar hardness values. However, extremely short exposure durations advocated by some manufacturers did not provide hardness values equivalent to those achieved with conventional treatment.

290

JOURNAL OF ESTHETIC DENTISTRY

Developments in Shrinkage Control of Adhesive Restoratives


BIBI S. DAUVILLIER, MSc* MAXIM P. AARNTS, PhD* ALBERT J. FEILZER, DDS, PhD*

ABSTRACT

Purpose: This article reviews material properties and application techniques important in minimizing effects of polymerization shrinkage during the curing reaction of resin composite restorative materials used in adhesive dentistry. Materials and Methods: Relevant scientific publications were critically reviewed. Results: Since it was recognized that shrinkage, which takes place during the curing reaction of resin composite restorative materials, may cause severe problems in adhesive dentistry, considerable effort has been put into reducing the negative effects. The most important problem is the debonding of the restorationtooth interface, resulting in increased microleakage and, ultimately, in secondary caries. Despite all efforts, there is still no material or general application method that guarantees a leak-proof and durable restoration.
CLINICAL SIGNIFICANCE

It is of the utmost importance that dental practitioners know how to deal with the problems related to resin composite shrinkage, so that they can choose the material and procedure most likely to produce a leak-proof and durable restoration, maximizing the potential for clinical success. (J Esthet Dent 12:291299, 2000)

esin composite restorative materials have gained a permanent position on the dental market. Their superior esthetics and consecutive preparation requirements (less destructive than amalgam) have been instrumental in this commercial success. The ideal restoration has a tight seal with remaining tooth structure, since otherwise, bacteria and toxins

produced by bacteria can invade and grow in the gap formed, resulting in pulp irritation and even secondary caries (Figure 1).1,2 This perfect adaptation must be obtained during curing and then maintained during thermal and mechanical cycling for the lifetime of the restoration or of the patient. Currently no commercially available resin composite guarantees an intact seal. Because the resin has no

anti-microbiologic activity, it is important that a restoration be placed in such a way that the best possible marginal seal is obtained. There are, however, many side effects that frustrate the goal of a perfectly sealed restoration. Most of these effects are related to polymerization shrinkage of the restoration during the curing process. Commercially available composites

*Department of Dental Materials Science, Academic Centre for Dentistry, Amsterdam, The Netherlands
VOLUME 12, NUMBER 6, 2000

291

DEVELOPMENTS IN SHRINKAGE CONTROL OF ADHESIVE RESTORATIVES

Figure 1. Effects of shrinkage stresses in restoration. (Reprinted from Versluis A, Tantbirojn D, Douglas WH. Do dental composites always shrink toward the light? J Dent Res 1998; 77:14351445, with permission.)

still undergo a volumetric shrinkage of 2 to 9%.36 Therefore, a major portion of this article is devoted to what the practitioner can do to minimize the negative effects of polymerization shrinkage.
SHRINKAGE

The matrix of most contemporary composite materials consists of methacrylate-based monomers.7 Volume reduction during curing results from closer packing of monomer molecules in the polymerized resin matrix.8 So, dimensional stability of the restoration is poor in the early stages of cure, whereas the density of the material increases. To prevent shrinkage, it is important to minimize the density difference between the cured and the uncured material.

Upon polymerization, unfilled resins containing mainly bis-GMA and TEGDMA undergo a volumetric shrinkage of approximately 9 to 10%.3 However, the presence of filler particles considerably reduces that shrinkage.9,10 At the same time, an increase in the percentage of filler loading is also accompanied by a significant drawback. The present generation of chemically or light-activated resin composites undergo a free volumetric shrinkage of 4 to 9% for flowable composites. For non-flowable or condensable composites, this value ranges from 2% to 5%, with most values near 3.5% of volume.46,9 Several variables are known to influence polymerization shrinkage. One variable is the size of the

monomer molecule undergoing polymerization. The larger the molecule before polymerization, the lower the polymerization shrinkage for a given volume of monomer.8,1113 Another influential variable is the volume fraction of the inorganic filler, including prepolymerized resin powder, within the composite. High filler loading results in lower polymerization shrinkage.9 This relation holds true until the point where a relatively high level of filler results in a clay-like paste, owing to increased viscosity. At high filler loading, the proportion of diluents (small monomers) in the resin system must increase to ensure acceptable handling properties. However, this addition may negate the effect of the high filler loading on cure shrinkage. Moreover, composite with high

292

JOURNAL OF ESTHETIC DENTISTRY

DAUVILLIER ET AL

filler loading results in a high degree of stiffness, which ultimately causes high shrinkage stress. Finally, the nature of the resin undergoing polymerization plays an important role in shrinkage. Several research groups are currently attempting to develop new resins that undergo less polymerization shrinkage.14,15 Commercial development of these resins may be many years away, as the process of gaining acceptance by the Food and Drug Administration (FDA) is time-consuming and expensive. However, if such resins ultimately are developed, they will largely eliminate the clinical consequences of polymerization shrinkage and will allow simple bulk placement of the material.
STRESS

In principle, a shrinking material pulls away from the weakest bond. In dental practice, the weakest bond is generally the free, unbonded surface of the restoration, provided that good adhesion between the restoration and the tooth is achieved. Adhesion to dentin is usually enhanced by the use of etching techniques, conditioners, bonding systems, and other means.16 Although of crucial importance, the subject of bonding systems is beyond the scope of this review, and in the remainder of the article an optimal adhesion between the tooth and restoration is assumed.

A large portion of shrinkage occurs in the early stages of the curing reaction: after about 15 minutes for chemically activated materials, and after about 60 seconds for light-activated materials (Figure 2). Thus, problems associated with adhesion loss often start during this early stage of cure, occasionally even before the patient has left the dentists chair. Voids or microcracks in the restoration or in enamel are formed during polymerization, when local stress exceeds polymer network strength. These voids and microcracks, as well as the poor interfa-

It should now be clear that shrinkage of resin composites, which up to now has been regarded as inevitable, must be controlled and directed toward the preparation walls, to prevent gap formation. However, as a result of adhesion to preparation walls, volumetric shrinkage is constrained. This constraint, in combination with an increasing modulus of elasticity, inevitably leads to development of stress. Although loss of adhesion from the tooth structure can occur at any time, the most likely moment is when the magnitude of shrinkage stress exceeds the strength of the developing restoration tooth bond.

Figure 2. Relation between axial shrinkage stress (y-axis) and axial shrinkage strain (x-axis) of a chemically activated resin composite (Silar, 3M Dental Products, St. Paul, Minnesota) and an analogous light-activated resin composite (Silux Plus, 3M Dental Products) during curing at room temperature for 1 hour. The chemically activated composite (C = 0.5) was mixed 1:1 w/w and the light-activated composite (C = 1.0) was exposed for 40 seconds with a light unit (Elipar Highlight, standard mode, ESPE, Seefeld, Germany) at distance of 4 mm. The light intensity at the light exit tip was 600 mW/cm2 (radiometer, model 100, Demetron Research Corp., Danbury, Connecticut). Note temperature effect after light exposure and difference in onset shrinkage strain.

VOLUME 12, NUMBER 6, 2000

293

DEVELOPMENTS IN SHRINKAGE CONTROL OF ADHESIVE RESTORATIVES

cial adhesion between filler and matrix, can encourage cohesive restoration fractures.17
STRESS RELIEF

Two factors have a major impact on the ultimate stress level of the restoration: the chemical and physical properties of a material and the way a material is handled during its application. The properties of materials are largely determined by the manufacturer, although a practitioner can influence those properties to some extent. For example, a dentist can alter the ratio of a twopaste system or use special curing lights and light-curing procedures that affect polymerization rate and degree of conversion.1820 Obviously, this manipulation will also influence the final material properties. The choice of a specific material, application method, or type of restoration can also have an impact on the ultimate quality of the restoration. This statement implies that practitioners must have considerable material expertise if they are to make a well-informed decision in favor of a particular material or application method. Chemical and Physical Properties For minimal impact on the integrity of a restored tooth, stress development must be minimized. One possible solution would be a reduction in the amount of polymerization shrinkage. Changing the chemical and physical properties of compos-

ite materials in such a way that shrinkage stress is no longer a problem is primarily the concern of those developing new resin composite materials. The development of nonshrinking materials (shrinkage lower than 0.4% of volume) might be a solution, but unfortunately there is no nonshrinking material on the market that can compete on all levels with conventional composite materials.15,2123 Moreover, the solution of one problem might very well create a new one (e.g., water sorption after curing might frustrate the high expectations of a nonshrinking material). Another approach to reduce shrinkage stress is to modify the resin composition so that the polymerization rate is lowered without influencing the degree of conversion. A slow polymerization rate may be expected to increase the ability of the material to flow without damaging its internal structure. In a restorative material with increased flow capacity, the volume change attributable to shrinkage is compensated by material flow from the unbonded, outer surface, ultimately resulting in lower stress. Resin composites can be chemically modified to reduce the polymerization rate in various ways. Use of less reactive resins is one possibility,24 but this method may have a negative effect on degree of conversion, resulting in more residual, unreacted monomer remaining in the cured

composite. Addition of retardative agents requires a careful choice of biocompatible chemicals.25 Reducing the amount of initiator system components requires no other chemicals than those already used in current systems.26,27 However, a balance must be found between a low reaction rate, on the one hand, and adequate conversion of the monomers, on the other hand. In all probability, the best way to obtain a lower polymerization rate in light-activated resin composites, together with a sufficiently high conversion, is by developing new initiator systems. Flow or Viscoelastic Behavior The solutions to stress reduction previously mentioned are mainly of interest to researchers and manufacturers of composites. However, the dentist has to deal with a wide variety of commercial products. Although materials differ in monomer composition, concentration of initiating system and filler type, size, loading, coating, resin composites can be divided into two general groups on the basis of polymerization rate: light-activated and chemically activated composites. Light-activated resin composites are popular among dentists because they can be cured on command. However, it has been demonstrated that, under the same conditions, light-activated composites generate higher polymerization shrinkage stress and more exothermic heat than

294

JOURNAL OF ESTHETIC DENTISTRY

DAUVILLIER ET AL

the analogous chemically activated composites.28,29 The dental literature has given considerable attention to a variety of methods designed specifically for light-activated resin composites, to reduce internal stresses in the restored tooth.30 During the early stages of polymerization, monomers are mainly converted into polymeric chains. After a certain degree of conversion has been attained, the predominant reaction is the cross-linking of the polymeric chains, resulting in a strong polymeric network.31 Although during the chain-growing period material viscosity rapidly increases, the polymeric chains can still slide along one another to relieve stress. This polymer chain movement is referred to as viscous flow behavior. When the cross-linking reaction becomes predominant,

there is no longer the ability of individual polymer chains to slide. At this stage, usually denoted as the post-gel phase, the polymeric chains reach sufficient modulus of elasticity to develop a strong, rigid viscoelastic material. Any further composite shrinkage will generate mechanical stress in the restoration. When adhesion survives the stress, microcracks or, in severe cases, voids can be generated in the viscoelastic material. Configuration Reduction of the negative effects of shrinkage stress can be controlled by practitioners; it involves the design of the preparation and the methods used to apply a restoration. The relation between the shape of a preparation and shrinkage stress development in composites has been demonstrated by Feilzer et al.32

In this context, the shape of the preparation is often described by means of the configuration factor (C-factor) (Figure 3). The C-factor denotes the ratio between the bonded and the free (unbonded) area of the restoration. It should be noted that the term bonded area means bonded to a rigid surface. In general, more bonded area leads to higher shrinkage stress, since composite flow is largely restricted to the small, free area of the material. This factor explains why the adhesive Class IV restoration has proven so successful, whereas other classes, in which the restoration is bordered by preparation walls (i.e., a high C-factor), often display marginal defects. The practitioner is in full control of preparation design; however, there are many other factors that influence the actual shape of the preparation, including some

Figure 3. A, The relation between different schematic, rectangular restorations, the corresponding configuration (C) factor values, and standard Class II, IV, and V restorations. B, Time-axial shrinkage stress relation of a chemically activated resin composite (Silar, 3M Dental Products, St. Paul, Minnesota) during curing at room temperature for various C-factor values. Reprinted from Rabek JF. Experimental and analytical methods for the investigation of radiation curing. In: Fouassier JP, Rabek JF, eds. Radiation curing in polymer science and technology. 1st Ed. London: Elsevier Science, 1993, with permission.

VOLUME 12, NUMBER 6, 2000

295

DEVELOPMENTS IN SHRINKAGE CONTROL OF ADHESIVE RESTORATIVES

with an undesirably high C-factor. There are several methods by which practitioners can reduce the impact of shrinkage stress on the quality of a restoration with a high C-factor. Layers, Liners, and Porosities Three methods described to reduce the effects of shrinkage stress are all based upon a reduction of the effective C-factor. For preparations with a large C-factor (Classes I and V), the dentist can apply a restorative material in several layers or increments. The advantage of this technique is twofold: (1) the C-factor for a small increment is lower than for bulk filling; and (2) small increment light-activated composites can be more thoroughly cured, since light intensity diminishes with the fourth power of light penetration.33 Of course, the main disadvantage of this method is that it is a timeconsuming procedure.34,35 It is thought that, when the walls of a preparation with an unfavorable (i.e., high) C-factor are covered with a relatively thick layer of a low elastic modulus material, the bulk shrinkage of the main restoration acquires some freedom of movement from the adhesive liner.36,37 This concept is feasible only when the liner extends to the cavosurface margin. Additionally, the elastic liner between the tooth and composite is often less wear-resistant at the restoration surface, resulting in surface pitting, which may provide a site for bacteria growth.38

The real effect of a low modulus lining material is probably its contribution to a more equal distribution of tensile and shear stresses over the adhesive interface. This material could dissipate the shear peak stress and generate no high polymerization shrinkage stress on the adhesive layer. Thus the adhesive, which is often not properly cured, owing to oxygen inhibition, is given time to cure before the high-bulk shrinkage stresses of the overlying, higher filler-loaded composite begin to act on it. Although the mechanisms are not clear, layering and low modulus liners are now generally accepted as a means of reducing polymerization shrinkage stress. Both methods have the disadvantage of additional time-consuming steps during restoration. The literature provides no clarity with respect to the sandwich technique, in which glass ionomer cements are used as liner.39 Deliberately admixing small air bubbles into a composite prior to use results in porosities in the cured composite. These porosities can be considered as unbonded areas, and they lead to a lower effective C-factor and, thus, lower shrinkage stress.28,40 However, it should be kept in mind that porosities can have a negative effect on other composite properties (e.g., water sorption, Youngs modulus, wear, etc.).4143 For this reason, the practice of deliberately inducing porosities in a composite should be discouraged.

Light Sources A recent method designed to reduce the polymerization rate of lightactivated resin-based materials involves varying light intensity on the restoration, either by reducing the output of the curing light or by increasing the distance between the light exit tip and the composite.18,44,45 A significant problem presented by the use of low light intensities is a reduced curing depth, which further declines when the quality of the light source in the curing unit deteriorates with age.4649 A predictive model for depth-of-cure devised by Rueggeberg et al suggests that the duration of curing compensates for the lower intensity.50 Although present-day conventional light-activated composites were developed for traditional procedures with a conventional halogen light source (4060 s exposure with light intensity 600 mW/cm2), many studies report significantly lower exposure durations involving new light units.19,44,5153 However, a valid comparison between light units requires that the spot diameter, intensity, wavelength distribution, exposure duration, and distance between light exit tip and composite must be specified. Failure to specify these parameters makes comparison among light units impossible. The physical and mechanical properties of composites are greatly influenced by the extent to which the resin has been cured.20,54,55 As curing efficiency and a lower polymerization rate may be

296

JOURNAL OF ESTHETIC DENTISTRY

DAUVILLIER ET AL

diametrically opposed to each other, a balance must be found between low shrinkage stress, on the one hand, and an adequate monomer conversion level, on the other.56 Hygroscopic Expansion The hygroscopic properties of a composite, although difficult to determine, can influence ultimate shrinkage stress.5763 Hygroscopic expansion (swelling) because of water sorption from saliva may, after curing, substantially relieve shrinkage stress.64 Unfortunately, swelling is much more marked for restorations with a low C-factor, in which shrinkage stress is not as great a problem. In the case of high C-factor restorations, the surface of the restoration, which is exposed to the oral cavity, will initially gain in volume. This gain produces a gradient from the outer surface to the bulk of the restoration, thus adding additional stress. Finally, owing to the slow process of water sorption from saliva, stress relief may come too late, after fractures have already formed. Although water sorption is generally recognized as a stress-relieving mechanism, there are few quantitative data available to assess its true impact. After a prolonged period of swelling, nonshrinking composite materials may encounter major problems related to expansion stress in some types of restorations (e.g., mesioocclusodistal restorations).

CONCLUSION

In the past 10 years, a great deal of effort has been made toward the development of nonshrinking and even expanding composite materials for dental applications. However, at present, the dental practitioner still has to deal with shrinking resin composites and the accompanying problems. Because there is as yet no easy, general solution to these problems, a proper understanding of the mechanisms causing the problems, and the methods that can be used to reduce their impact on the quality of a restoration, is of crucial importance. The information presented is intended to help the clinician obtain maximum benefit from the selection and application of resin composites.
DISCLOSURE AND ACKNOWLEDGMENT

5. Feilzer AJ, De Gee AJ, Davidson CL. Curing contraction of composites and glass ionomer cements. J Prosthet Dent 1988; 59:297300. 6. Watts DC, Cash AJ. Determination of polymerization shrinkage kinetics in visible light-cured materials: methods development. Dent Mater 1991; 7:281287. 7. Deb S. Polymers in dentistry. Proc Inst Mech Eng [H] 1998; 212:453464. 8. Venhoven BA, de Gee AJ, Davidson CL. Polymerization contraction and conversion of light-curing BisGMA-based methacrylate resins. Biomaterials 1993; 14:871875. 9. Miyazaki M, Hinoura K, Onose H, Moore BK. Effect of filler content of light-cured composites on bond strength to bovine dentine. J Dent 1991; 19:301303. 10. Munksgaard EC, Hansen EK, Kato H. Wall-to-wall polymerization contraction of composite resins versus filler content. Scand J Dent Res 1987; 95:526531. 11. Culbertson BM, Wan QC, Tong YH. Preparation and evaluation of visible lightcured multi-methacrylates for dental composites. J Macromol Sci Pure Appl Chem 1997; A34:24052421. 12. Davy KW, Kalachandra S, Pandain MS, Braden M. Relationship between composite matrix molecular structure and properties. Biomaterials 1998; 19:20072014. 13. Nie J, Linden LA, Rabek JF, Ekstrand J. Photocuring of mono- and di-functional (meth)acrylates with tris [2-(acryloyloxy) ethyl]isocyanurate. Eur Polymer J 1999; 35:14911500. 14. Stansbury JW, Dickens B, Liu DW. Preparation and characterization of cyclopolymerizable resin formulations. J Dent Res 1995; 74:11101115. 15. Eick JD, Robinson SJ, Byerley TJ, Chappelow CC. Adhesives and nonshrinking dental resins of the future. Quintessence Int 1993; 24:632640. 16. Eick JD, Gwinnett AJ, Pashley DH, Robinson SJ. Current concepts on adhesion to dentin. Crit Rev Oral Biol Med 1997; 8:306335. 17. Kim KH, Park JH, Imai Y, Kishi T. Microfracture mechanisms of dental resin composites containing spherically-shaped filler particles. J Dent Res 1994; 73:499504.

The authors thank Mrs. B. Fasting for her comments on the English text.
REFERENCES

1. Brannstrom M, Vojinovic O. Response of the dental pulp to invasion of bacteria around three filling materials. ASDC J Dent Child 1976; 43:8389. 2. Jorgensen KD, Asmussen E, Shimokobe H. Enamel damages caused by contracting restorative resins. Scand J Dent Res 1975; 83:120122. 3. Labella R, Lambrechts P, Van Meerbeek B, Vanherle G. Polymerization shrinkage and elasticity of flowable composites and filled adhesives. Dent Mater 1999; 15:128137. 4. Feilzer AJ, Dooren LH, De Gee AJ, Davidson CL. Influence of light intensity on polymerization shrinkage and integrity of restoration-cavity interface. Eur J Oral Sci 1995; 103:322326.

VOLUME 12, NUMBER 6, 2000

297

DEVELOPMENTS IN SHRINKAGE CONTROL OF ADHESIVE RESTORATIVES

18. Silikas N, Eliades G, Watts DC. Light intensity effects on resin-composite degree of conversion and shrinkage strain. Dent Mater 2000; 16:292296. 19. Munksgaard EC, Peutzfeldt A, Asmussen E. Elution of TEGDMA and BisGMA from a resin and a resin composite cured with halogen or plasma light. Eur J Oral Sci 2000; 108:341345. 20. Peutzfeldt A, Sahafi A, Asmussen E. Characterization of resin composites polymerized with plasma arc curing units. Dent Mater 2000; 16:330336. 21. Stansbury JW. Cyclopolymerizable monomers for use in dental resin composites. J Dent Res 1990; 69:844848. 22. Stansbury JW. Synthesis and evaluation of new oxaspiro monomers for double ringopening polymerization. J Dent Res 1992; 71:14081412. 23. Byerley TJ, Eick JD, Chen GP, Chappelow CC, Millich F. Synthesis and polymerization of new expanding dental monomers. Dent Mater 1992; 8:345350. 24. Labella R, Davy KW, Lambrechts P, Van Meerbeek B, Vanherle G. Monomethacrylate co-monomers for dental resins. Eur J Oral Sci 1998; 106:816824. 25. DeCaprio AP. The toxicology of hydroquinone: relevance to occupational and environmental exposure. Crit Rev Toxicol 1999; 29:283330. 26. Kalliyana Krishnan V, Yamuna V. Effect of initiator concentration, exposure time, and particle size of the filler upon the mechanical properties of a light-curing radiopaque dental composite. J Oral Rehabil 1998; 25:747751. 27. Cook WD. Photopolymerization kinetics of dimethacrylates using the camphorquinone amine initiator system. Polymer 1992; 33:600609. 28. Feilzer AJ, de Gee AJ, Davidson CL. Setting stresses in composites for two different curing modes. Dent Mater 1993; 9:25. 29. Kinomoto Y, Torii M, Takeshige F, Ebisu S. Comparison of polymerization contraction stresses between self- and light-curing composites. J Dent 1999; 27:383389. 30. Bouschlicher MR, Vargas MA, Boyer DB. Effect of composite type, light intensity,

configuration factor and laser polymerization on polymerization contraction forces. Am J Dent 1997; 10:8896. 31. Rabek JF. Experimental and analytical methods for the investigation of radiation curing. In: Fouassier JP, Rabek JF, eds. Radiation curing in polymer science and technology. 1st Ed. London: Elsevier Science, 1993. 32. Feilzer AJ, De Gee AJ, Davidson CL. Setting stress in composite resin in relation to configuration of the restoration. J Dent Res 1987; 66:16361639. 33. Skoog DA. Raman spectroscopy. In: Principles of instrumental analysis. 3rd Ed. New York: CBS College Publishing, 1985:362. 34. Crim GA. Microleakage of three resin placement techniques. Am J Dent 1991; 4:6972. 35. Lutz E, Krejci I, Oldenburg TR. Elimination of polymerization stresses at the margins of posterior composite resin restorations: a new restorative technique. Quintessence Int 1986; 17:777784. 36. Kemp-Scholte CM, Davidson CL. Marginal integrity related to bond strength and strain capacity of composite resin restorative systems. J Prosthet Dent 1990; 64:658664. 37. Kemp-Scholte CM, Davidson CL. Complete marginal seal of Class V resin composite restorations effected by increased flexibility. J Dent Res 1990; 69:12401243. 38. Benderli Y, Ulukapi H, Balkanli O, Kulekci G. In vitro plaque formation on some dental filling materials. J Oral Rehabil 1997; 24:8083. 39. Woolford M. Composite resin attached to glass polyalkenoate (ionomer) cement: the laminate technique. J Dent 1993; 21:3138. 40. Alster D, Feilzer AJ, De Gee AJ, Mol A, Davidson CL. The dependence of shrinkage stress reduction on porosity concentration in thin resin layers. J Dent Res 1992; 71:16191622. (Published erratum appears in J Dent Res 1993; 72:87.) 41. Barkmeier WW, Erickson RL. Shear bond strength of composite to enamel and dentin using Scotchbond Multi-Purpose. Am J Dent 1994; 7:175179.

42. Jorgensen KD, Hisamitsu H. Porosity in microfill restorative composites cured by visible light. Scand J Dent Res 1983; 91:396405. 43. McCabe JF, Ogden AR. The relationship between porosity, compressive fatigue, limit and wear in composite resin restorative materials. Dent Mater 1987; 3:912. 44. Unterbrink GL, Muessner R. Influence of light intensity on two restorative systems. J Dent 1995; 23:183189. 45. Uno S, Asmussen E. Marginal adaptation of a restorative resin polymerized at reduced rate. Scand J Dent Res 1991; 99:440444. 46. Rueggeberg FA, Caughman WF. The influence of light exposure on polymerization of dual-cure resin cements. Oper Dent 1993; 18:4855. 47. Rueggeberg FA, Caughman WF, Curtis JW Jr., Davis HC. Factors affecting cure at depths within light-activated resin composites. Am J Dent 1993; 6:9195. 48. Miyazaki M, Hattori T, Ichiishi Y, Kondo M, Onose H, Moore BK. Evaluation of curing units used in private dental offices. Oper Dent 1998; 23:5054. 49. Pilo R, Oelgiesser D, Cardash HS. A survey of output intensity and potential for depth of cure among light-curing units in clinical use. J Dent 1999; 27:235241. 50. Rueggeberg FA, Caughman WF, Curtis JW Jr, Davis HC. A predictive model for the polymerization of photo-activated resin composites. Int J Prosthodont 1994; 7:159166. 51. Rueggeberg FA, Caughman WF, Chan DC. Novel approach to measure composite conversion kinetics during exposure with stepped or continuous light-curing. J Esthet Dent 1999; 11:197205. 52. Solomon CS, Osman YI. Evaluating the efficacy of curing lights. SADJ 1999; 54:357362. 53. Burgess JO, DeGoes M, Walker R, Ripps AH. An evaluation of four light-curing units comparing soft and hard curing. Pract Periodont Aesthet Dent 1999; 11:125132. 54. Ruyter IE. Methacrylate-based polymeric dental materials: conversion and related properties. Summary and review. Acta Odontol Scand 1982; 40:359376.

298

JOURNAL OF ESTHETIC DENTISTRY

DAUVILLIER ET AL

55. Asmussen E. Factors affecting the quantity of remaining double bonds in restorative resin polymers. Scand J Dent Res 1982; 90:490496. 56. Ferracane JL, Greener EH. The effect of resin formulation on the degree of conversion and mechanical properties of dental restorative resins. J Biomed Mater Res 1986; 20:121131. 57. Pearson GJ. Long-term water sorption and solubility of composite filling materials. J Dent 1979; 7:6468. 58. Hansen EK. Visible light-cured composite resins: polymerization contraction, contraction pattern and hygroscopic expansion. Scand J Dent Res 1982; 90:329335.

59. Bowen RL, Rapson JE, Dickson G. Hardening shrinkage and hygroscopic expansion of composite resins. J Dent Res 1982; 61:654658. 60. Feilzer AJ, Kakaboura AI, de Gee AJ, Davidson CL. The influence of water sorption on the development of setting shrinkage stress in traditional and resin-modified glass ionomer cements. Dent Mater 1995; 11:186190. 61. Fan PL, Edahl A, Leung RL, Stanford JW. Alternative interpretations of water sorption values of composite resins. J Dent Res 1985; 64:7880. 62. Hirasawa T, Hirano S, Hirabayashi S, Harashima I, Aizawa M. Initial dimensional change of composites in dry and wet conditions. J Dent Res 1983; 62:2831.

63. Soderholm KJ. Water sorption in a bis(GMA)/TEGDMA resin. J Biomed Mater Res 1984; 18:271279. 64. Feilzer AJ, de Gee AJ, Davidson CL. Relaxation of polymerization contraction shear stress by hygroscopic expansion. J Dent Res 1990; 69:3639. 65. Versluis A, Tantbirojn D, Douglas WH. Do dental composites always shrink toward the light? J Dent Res 1998; 77:14351445. (Comments)

Reprint address: Albert J. Feilzer, Department of Dental Material Science, Academic Centre for Dentistry Amsterdam, Louwesweg 1, NL-1066 EA Amsterdam, The Netherlands; e-mail: a.feilzer@acta.nl 2000 BC Decker Inc

VOLUME 12, NUMBER 6, 2000

299

Curing Dental Resins and Composites by Photopolymerization


JEFFREY W. STANSBURY, PHD*

ABSTRACT

The development and continued evolution of photopolymerizable dental materials, particularly dental composite restoratives, represent a significant, practical advance for dentistry. The highly successful integration of the light-activated curing process for dental applications is described in this review. The basic mechanisms by which the photoinitiators efficiently convert monomers into polymers are discussed along with the variety of factors that influence the photopolymerization process. The conventional camphorquinone-amine visible light photoinitiator system used in most dental restorative materials is illustrated in addition to some alternative initiator systems that have been studied for dental materials applications.
CLINICAL SIGNIFICANCE

Photopolymerization has become an integral component of the practice of dentistry. A better appreciation of the photopolymerization process as well as its potential and limitations may aid the dentist in the delivery of both esthetic and restorative dental care. ( J Esthet Dent 12:300308, 2000)

he introduction of photopolymerization to dentistry began nearly coincident with the commercialization of this technology in the late 1960s. Initially, ultraviolet (UV)-cured pit and fissure sealants were put into clinical practice,1 and soon thereafter, adhesive and restorative applications of lightcured materials followed.2 Since that time, photoactivated dental materials have stayed current with and contributed to state-of-the-art advancements in photopolymerization technology. A variety of photocurable dental products, such as

sealants, adhesives, dentures, and impression materials, have been developed. However, the primary application of this technique is seen in composite restoratives. Nuva Fil (Dentsply, York, Pennsylvania) was introduced in 1972. It did not take long for dental practitioners to accept the modification to chemically cured composites that were based on a two-paste benzoyl peroxide-amine initiator system. This new technology offered significant advantages: (1) a single-paste system requiring no mixing and (2) complete operator control over working time with

a rapid cure commencing on command. Along with these benefits came the need to develop new restorative placement techniques, mainly the use of multiple (incremental) layers to place large restorations, since the photopolymerization of early generation composites had serious limitations with respect to depth of cure.3 Also related to curing efficiency, the shade of composite material being placed as well as the day-to-day variation in output intensity of the curing lamp were other clinically relevant concerns raised with the

*University of Colorado Health Sciences Center, School of Dentistry, Aurora, Colorado 300
JOURNAL OF ESTHETIC DENTISTRY

STANSBURY

advent of photopolymerization in dental practice. By the 1980s, the original UV-curable materials, based on benzoin methyl ether or other UV active photoinitiators, were reformulated to incorporate new visible light wavelength initiators. The uniform change to visible light initiation within the dental materials industry avoided problems of UV-promoted tissue damage, as well as concerns of ophthalmologic damage to practitioners.4,5
PHOTOINITIATION

Figure 1. UV-induced alpha-cleavage of benzoin methyl ether (BME).

A photoinitiator is a molecule that can absorb light and, as a result, either directly or indirectly, generate a reactive species that can then initiate polymerization.6 Initiators generally have a carbonyl group with nonbonding electrons that can be promoted to a * antibonding orbital (lower energy bonding and higher energy antibonding molecular orbitals are created by the respective positive and negative overlap of atomic orbitals) by absorption of light at the appropriate wavelength. With initiators like benzoin methyl ether (Figure 1), the result of exposure to light is intramolecular -cleavage (carbon carbon bond fragmentation at the carbonyl group) to yield two free radicals, both of which have the potential to initiate polymerization. In current dental resins, camphorquinone (CQ) is typically used as a visible light-activated free radical photoinitiator (Figure 2). Cam-

phorquinone is an example of a photoinitiator that requires a coinitiator for an efficient polymerization process to occur. A co-initiator is a separate compound that does not absorb light but interacts with an activated photoinitiator to produce a reactive species. In the case of a dental composite restorative containing CQ, a tertiary amine photoreductant is used as the coinitiator to provide the reactive radicals that begin polymerization. In some initiator systems, a photosensitizer is also included to absorb light at one wavelength and transfer the energy from its excited state to a photoinitiator that absorbs in a different wavelength range. Exam-

ples of this energy transfer are found in the visible light polymerizations conducted with cationic initiators that typically absorb in the UV region but can be sensitized to function at longer wavelengths by the inclusion of trace amounts of dyes or other active compounds including CQ.79 The efficiency of any given photoinitiator is governed by a number of factors. For high absorptivity to be achieved there obviously must be a relatively good match between the absorption spectrum of the photoinitiator and the emission spectrum of the light source. The absorption spectrum of CQ has a broad maxi-

Figure 2. Structure of camphorquinone (CQ) and visible light-activated free radical generation pathway.

VOLUME 12, NUMBER 6, 2000

301

CURING DENTAL RESINS AND COMPOSITES BY PHOTOPOLYMERIZATION

mum at 468 nm in the blue region of the visible spectrum. Dental curing units are generally halogen sources with bandpass filters that transmit in the 400- to 540-nm visible region. This filtering minimizes the harmful potential of either infrared radiation, which can give rise to a significant amount of heat, or UV radiation that can cause biologic damage. A useful photoinitiator must also have a high molar extinction coefficient (high absorptivity at low concentration). The compound should also undergo high yield intersystem crossing in the excited state (a change from the initially formed unreactive electron-promoted singlet state to the reactive triplet state) that can lead to productive chemical processes. These factors, as well as competitive deactivation pathways, such as initiator quenching by monomer or oxygen,10 control the number of active radical species produced per photon of light absorbed. The wavelength, intensity, and exposure duration of the incident light define the irradiation energy. These factors, along with choice of the photoinitiator concentration, permit a high degree of external control over the photopolymerization process. Beyond this consideration, the resin composition, which includes individual monomer structures and comonomer ratios, also exerts considerable influence on the polymerization as network formation proceeds.1113

CAMPHORQUINONE-AMINE PHOTOINITIATION SYSTEMS

Traditionally, CQ-tertiary amine initiators have been the standard in dental composite restoratives, and as such, a number of studies have been undertaken to determine the mechanism of initiation and the parameters that affect photopolymerization. Thorough studies of the CQ-amine visible light photoinitiation process and its influence on polymerization kinetics have been described.14,15 With a relatively broad absorption spectrum in the visible region (400550 nm, max = 468 nm), radiation in this range promotes an electron in one of the two carbonyl groups of CQ to a short-lived, excited energy state (half-life of approximately 0.05 ms for the CQ triplet). If, prior to its decay or deactivation, it encounters an amine molecule through diffusion or a preexisting association, an exciplex (excited state complex) can form (see Figure 2). Within this exciplex, the amine can donate to CQ first an electron to form the radical ion pair and then a proton to generate the free radical species. It is the aminoalkyl radical that initiates monomer polymerization, whereas the CQ-based counter radical may actually retard polymerization through termination reactions with growing polymer chains. The chain reaction process that defines free radical addition polymerization means that the absorption of a single photon of light by the initiator

can result in the incorporation of hundreds of monomer units into the polymeric network. Studies of the CQ-amine photoinitiation process have generally involved either evaluation of polymerization kinetics during photopolymerization or measurement of the degree of conversion obtained following polymerization. These polymerization-based measurement techniques, predominantly with infrared spectroscopy, directly determine conversion.16,17 Calorimetric methods that allow conversion to be calculated based on the heat released by the polymerization reaction have also been used.1820 Other, more indirect approaches involve analysis of some physical or mechanical property of the resulting polymer, such as flexural strength or microhardness.21 It has been amply demonstrated that properties of the restorative materials tend to improve as the level of conversion attained during photopolymerization is increased.22 Direct observation of the photoinitiation step, rather than the resulting polymerization process, has been accomplished with time-resolved laser spectroscopy, steady-state UVvisible light spectroscopy, nuclear magnetic resonance spectroscopy, and electron paramagnetic resonance (EPR) techniques.6 These direct observation approaches, as well as comparative photopolymerization

302

JOURNAL OF ESTHETIC DENTISTRY

STANSBURY

studies, show that the structure of the amine photoreductant used with CQ has a dramatic effect on the efficacy of the initiating system. A wide variety of aromatic and aliphatic amines have been compared for their initiating potential with CQ.14,23 Aromatic tertiary amines were found to be somewhat more reactive co-initiators than the aliphatic counterparts.15 Work with chemically cured initiator systems (benzoyl peroxide-amine) demonstrated that aromatic amines were clearly more effective than aliphatic amines.24 However, tertiary aliphatic amines, predominantly the copolymerizable N,N-dimethylaminoethyl methacrylate, are widely used in the photoinitiator systems of commercial dental restorative materials. Conversely, primary amines, which bear active hydrogens, function as inhibitors of the photopolymerization process. The effect of the CQ:amine ratio on initiator efficiency has been examined by the measurement of conversion in unfilled resins. At a fixed CQ concentration, it was found that conversion increased monotonically to approximately a 1:2 or 1:3 molar ratio of CQ to amine and then plateaued with additional amine.25 A separate study showed that CQ levels at mole fractions between approximately 0.5% and 1.0% provided full conversion, whereas higher CQ concentrations gave no additional improvements and yielded discolored polymer specimens.26,27

PHOTOPOLYMERIZATION IN DENTISTRY

The use of EPR techniques to study photopolymerization is particularly appropriate since they provide methods to directly monitor free radical population throughout the polymerization process.28 During the initial stages, radical concentration increases rapidly and reaches an essentially steady-state condition where new radical production is balanced by radical termination processes. In one such investigation, the irradiation time necessary to achieve maximum radical formation was determined as a function of specimen thickness.29 An exponential increase in irradiation time was necessary to ensure maximal conversion as composite thickness increased. Using a standard dental curing unit, a 2-mm-thick composite specimen required approximately 100 seconds to reach a maximum radical concentration. Other EPR-based studies, which also concluded that typically used exposure durations are not sufficient to achieve maximum radical production, demonstrated the long-term persistence of radicals that become trapped in the glassy polymeric matrix.30,31 Lifetimes of these trapped radicals are directly related to the stiffness of the polymeric network and can vary from hours to months. The fillers used in composite restoratives appear to significantly enhance the decay of trapped radicals. Diffusion of oxygen

through the polymer can result in conversion of the highly reactive carbon-based radicals to relatively stable hydroperoxides, which in sufficient concentrations can alter bulk properties of the polymer.12 Diffusion of oxygen into the exposed resin or composite surface as polymerization proceeds results in quenching of both initiator and polymer-based radical species and is responsible for the poorly polymerized, air-inhibited surface layer.32 Another consequence of radicals trapped within polymers is the protracted post-cure, or additional conversion, that continues after irradiation is ended. The majority of potential conversion that can be attained with a given resin composition is achieved during the first few minutes after irradiation; however, significant increases in both conversion and, particularly, the evolution of polymer properties are observed up to and beyond 24 hours.21 This post-cure process can be greatly facilitated by the application of heat, which increases mobility within the polymeric network and allows free monomer and pendant chains to encounter remaining radical sites and react further.22 Thus, resin-based materials for indirect restorative applications are typically photocured at elevated temperatures or are given a postirradiation heat treatment to achieve maximum conversion and physical properties. The effect of delaying

VOLUME 12, NUMBER 6, 2000

303

CURING DENTAL RESINS AND COMPOSITES BY PHOTOPOLYMERIZATION

the time between the irradiation step and the post-cure heating has been studied as a demonstration of the effect of trapped radicals on the continuation of the curing process.33,34 It has also been shown that resins with higher concentrations of low viscosity diluent monomers, such as triethylene glycol dimethacrylate (TEGDMA), reach higher immediate degrees of conversion, but have less additional conversion during the post-cure interval than do resins that contain less diluent co-monomer.35,36 For direct filling materials, the temperature differential generated during photopolymerization can become a clinical concern. Factors, including the wavelength, intensity, and exposure duration as well as restoration composition and geometry, affect the temperature rise that can be expected. The vinyl addition polymerization reaction is exothermic by approximately 55 kJ/mol, based on the methacrylate functional groups.19 Compounding this effect, certain curing lights can impart a significant thermal rise, owing simply to the absorbed photon energy. Inclusion of substantial amounts of inert filler in composites essentially mitigates the potential exothermic response by diluting the reactive group density of the resin and serving as a heat sink. A recent study involving conventional dental curing lights has demonstrated temperature deviations at the cavity floor between 3C and

7C during photocuring of 2-mm thick commercial composite specimens.37 It appeared that a majority of the temperature elevation observed was attributable to direct energy input from the light sources used and not from the exothermic heat of the polymerization reaction. The photopolymerization process in dental composites is complicated by the presence of inorganic fillers of various particle sizes, which can range in from approximately 0.04 m to 10 m depending on the material.38 A reasonably close refractive index match between the resin matrix and the filler is an important factor to achieve efficient light throughput as well as yielding esthetically pleasing translucent polymerized composites.39,40 Since refractive index of the resin shifts to a significantly higher value as the polymerization occurs, light transmission is affected.41 Light interacts with filler particles, resulting in absorption and scattering that significantly attenuates the irradiation intensity reaching deeper portions of the composite. Microfilled resins allow less light penetration than hybrid and small particle-filled products.42 The effects of variation in irradiation wavelength and exposure duration on the polymerization efficiency of CQ-amine-activated resin composites have been investigated.43 Over the range of intensities associated with conventional dental curing lamps, a spectral output that

overlaps a significant portion of the CQ absorption range was found to be more efficient than a narrow irradiation band focused at the wavelength of the CQ absorption maximum. Comparisons of photopolymerization conversion achieved with equal light flux but varied intensities (short-duration exposure at high intensity compared with longduration exposure at low intensity) demonstrated that modest to moderate increases in conversion were associated with higher light intensities.44 Other studies have focused on the depth of cure of composites or on the significant differences in physical properties noted as a function of depth.45 In nearly every instance, even with extended irradiation intervals, a discrepancy is observed between properties, such as microhardness, measured at the upper and lower surfaces of typical laboratory specimens. In addition, absorption of light of the effective wavelength by the initiator in the upper region of the composite acts as a filter for the light being transmitted to the lower boundary of the material. Therefore, delays of greater than 40 seconds have been observed between the onset of irradiation and the transmission of light of the active wavelength from a 2-mm-thick composite specimen.46
ALTERNATIVE PHOTOINITIATOR SYSTEMS

Variations in visible light photoinitiator formulations for use in dental materials have been introduced in

304

JOURNAL OF ESTHETIC DENTISTRY

STANSBURY

efforts to enhance photopolymerization efficiency. New initiators may potentially provide higher degrees of conversion or faster cure rates with minimal light exposure and reduced initiator concentration. A simple alteration that can be made is addition of peroxide co-initiators to the CQ photoinitiator.47 In contrast to the increased cure rates noted with additives such as di-t-butyl peroxide, addition of more stable hydroperoxide components in the experimental photoinitiator system decreased polymerization rates. As another modification to the conventional CQ-amine initiator system, addition of 1-phenyl-1,2-propanedione (PPD) has been examined (Figure 3).48 Because the absorption maximum of PPD occurs at 410 nm, the polymers produced were less yellow than those prepared with only CQ as the initiator. A decrease in initiator-related discoloration after photocuring is clinically important in obtaining and maintaining an acceptable match of a restoration with adjacent tooth structure. The synergistic combination of PPD with CQ allows higher degrees of conversion to be achieved compared with only the CQ-amine initiator. The observed benefits may be attributable to the generation of free radicals from two distinctly different initiation mechanisms: proton abstraction with CQ-amine and direct intramolecular cleavage with PPD. Extensive studies were conducted on the use of aldehydes or diketones as reactive

Figure 3. Structures of aldehyde and diketone additives to the photoinitiator systems. PPD = l-phenyl-1,2-propanedione; PA = propionaldehyde; BD = butanedione.

ingredients in photoinitiated dental resins.49 Thus, substantial proportions of propionaldehyde (PA) and 2,3-butanedione (BD) additives enhanced conversion and improved mechanical properties with respect to the baseline materials with a chaintransfer mechanism proposed to account for the results (see Figure 3). Other studies evaluated the potential of photoinitiator systems that do not rely on camphorquinone as their basis. An example of this type of system is the investigation of a series of bisacylphosphine oxide (BAP) initiators (Figure 4).50 The various initiator structures examined all have absorption maxima in the UV region from 320 nm to 390 nm; however, the shoulder of these maxima extends well into the visible light wavelength and allows the ini-

tiators to be used with conventional visible light dental curing units. The initiation mechanism is a direct intramolecular cleavage that requires no amine co-initiator. A rapid, light-induced disappearance of initiator-imparted color, or photobleaching, that occurs during photopolymerization with initiators of this type allows thick polymer samples to be cured with an esthetic appearance. Similar initiators, triacylphosphine oxides (TPO) are also being examined as a co-initiator with CQ (see Figure 4). More complex three-component initiator systems are also currently being developed for improved polymerization efficiency and decreased sensitivity toward oxygen inhibition.51 In other cases, the photoinitiator composition has been varied to

Figure 4. Structures of phosphine oxide-based photoinitiators. BAP = bisacylphosphine oxide; TPO = triacylphosphine oxide.

VOLUME 12, NUMBER 6, 2000

305

CURING DENTAL RESINS AND COMPOSITES BY PHOTOPOLYMERIZATION

accommodate the demands imposed by specific dental materials applications. In dentin bonding, which relies on infiltration and polymerization of monomers within demineralized dentin, a water soluble photoinitiator, 2-hydroxy-3-(3,4-dimethyl-9oxo-9H-thioxanthen-2-yloxy)N,N,N-trimethyl-1-propanaminium chloride (QTX) has demonstrated potential to improve bond strengths (Figure 5).52,53 With an absorption maxium at 402 nm, QTX used in the aqueous primer solution applied to dentin replaces CQ in the role of providing the critical interfacial polymerization of the subsequent bonding resin and overlying composite layers.54 The development of new resin systems for dental restorative applications can also necessitate the introduction of new photoinitiators. Recent work directed toward a novel resin system based on epoxy rather than methacrylate curing chemistry has required use of cationic photoinitiators, such as diaryl iodonium hexafluoroantimonates (DIH) (see Figure 5).9,55 These initiators typically

contain an aliphatic side-chain on one of the aromatic rings, designed to improve solubility in the resin. The iodonium photocationic initiators are active in the UV region, and thus, CQ has been added to serve as a visible wavelength sensitizer. A novel approach to the study of the dental resin photopolymerization process involves the use of an iniferter, or living free radical initiator.56 p-Xylene bis(N,N-diethyl dithiocarbamate) (XDT) was used in the UV-initiated photopolymerization to yield model cross-linked acrylic polymeric networks (see Figure 5). Photocleavage of XDT occurs to yield a carbon-based initiating radical and a sulfur-based dithiocarbamate radical that remains associated with the terminus of the growing polymer chain. Although the polymerization is relatively slow, the polymer formed does not undergo bimolecular radical termination reactions or have trapped radicals in the network. This allows characterization of the resulting polymers by thermal

analysis techniques without causing an additional post-curing reaction that invariably alters polymer properties under observation.
CONCLUSION

The use of photopolymerization offers the dental practitioner a remarkable degree of control over the polymerization process. This is evidenced by the proliferation of light-curing units on the market. Whereas the conventional curing lights have tended toward higher power density outputs, other curing units now provide low intensity initial irradiation followed by a ramping up to a higher intensity level. Other devices offer the delivery of extremely intense doses of irradiation designed for rapid cure. Along with the availability of these and other photocuring technologies, dental materials researchers and manufacturers will undoubtedly continue to improve both the photoinitiator systems and the resins used in restorative materials. The benefits of more efficient resin polymerization may be expressed as further enhancements in restorative performance and durability and improved biocompatibility and color stability.
DISCLOSURE

Figure 5. Structures of application-specific photoinitiators. QTX = 2-hydroxy-3-(3,4 dimethyl-9-oxo-9H-thioxanthen-2-yloxy)-N,N,Ntrimethyl-1-propanaminium chloride; DIH = diaryl iodonium hexafluoroantimonate.

The author has no financial interest in any of the companies or products mentioned in his article.

306

JOURNAL OF ESTHETIC DENTISTRY

STANSBURY

REFERENCES

1. Buonocore MG. Adhesive sealing of pits and fissures for carries prevention, with use of ultraviolet light. J Am Dent Assoc 1970; 80:324328. 2. Rock WP. The use of UV radiation in dentistry. Br Dent J 1974; 136:455458. 3. Cook WD. Factors affecting the depth of cure of UV-polymerized composites. J Dent Res 1980; 59:800808. 4. Gange RW, Rosen CF. UVA effects on mammalian skin and cells. Photochem Photobiol 1986; 43:701705. 5. Ham WT, Mueller HA, Sliney DH. Retinal sensitivity to damage from short wavelength light. Nature 1976; 260:153155. 6. Fouassier JP. Photoinitiation, photopolymerization and photocuring: fundamentals and applications. Munich: Carl Hanser Verlag, 1995. 7. Crivello JV, Lam JHW. Dye-sensitized photoinitiated cationic polymerization. J Polym Sci Polym Chem Ed 1978; 16:24412451. 8. Bi Y, Neckers DC. A visible light initiating system for free radical promoted cationic polymerization. Macromolecules 1994; 27:36833693. 9. Chappelow CC, Pinzino CS, Jeang L, Harris CD, Holder AJ, Eick JD. Photoreactivity of expanding monomers and epoxybased matrix resin systems. J Appl Polym Sci 2000; 76:17151724. 10. Rueggeberg FA, Margeson DH. The effect of oxygen inhibition on an unfilled/filled composite system. J Dent Res 1990; 69:16521658. 11. Reed BB, Choi KM, Dickens SH, Stansbury JW. Effect of resin composition on kinetics of dimethacrylate photopolymerization. Polym Prepr (Am Chem Soc) 1997; 38:108109. 12. Lovell LG, Berchtold KA, Elliott JE, Lu H, Bowman CN. Understanding the kinetics and network formation of dimethacrylate dental resins. Polym Adv Technol. (Accepted). 13. Stansbury JW, Dickens SH. Network formation and compositional drift during photoinitiated copolymerization of dimethacrylate monomers. Polymer. (Submitted)

14. Cook WD. Photopolymerization kinetics of dimethacrylates using the camphorquinone amine initiator system. Polymer 1992; 33:600609. 15. Mateo JL, Bosch P, Lozano AE. Reactivity of radicals derived from dimethylanilines in acrylic photopolymerization. Macromolecules 1994; 27:77947799. 16. Ruyter IE, Gyorosi PP. Infrared spectroscopic study of sealants. Scand J Dent Res 1976; 84:396400. 17. Ruyter IE, Svendsen SA. Remaining methacrylate groups in composite restorative materials. Acta Odontol Scand 1978; 36:7582. 18. Antonucci JM, Toth EE. Extent of polymerization of dental resins by differential scanning calorimetry. J Dent Res 1983; 23:704707. 19. Miyazaki K, Horibe T. Polymerization of multifunctional methacrylates and acrylates. J Biomed Mater Res 1988; 22:10111022. 20. Urabe H, Wakasa K, Yamaki M. Cure performance of multifunctional monomers to photo-initiators: a thermoanalytical study on bis-GMA-based resins. J Mater Sci 1991; 26:31853190. 21. Leung RL, Fan PL, Johnston WM. Postirradiation polymerization of visible-light activated composite resin. J Dent Res 1983; 62:363365. 22. Ferracane JL, Condon JR. Postcure heattreatments for composites: properties and fractography. Dent Mater 1992; 8:290295. 23. Kerby RE, Tiba A, Culbertson BM, Schricker S, Knobloch L. Evaluation of tertiary amine co-intiators using differential scanning photocalorimetry. J Macromol Sci, Pure Appl Chem 1999; 36:12271239. 24. Antonucci JM, Peckoo RJ, Schruhl C, Toth EE. Slow-acting amine polymerization accelerators: para-dimethylaminobenzoic acid and its ethyl ester. J Dent Res 1981; 60:13251331. 25. Yoshida K, Greener EH. Effects of two amine reducing agents on the degree of conversion and physical properties of an unfilled light-cured resin. Dent Mater 1993; 9:246251.

26. Venhoven BAM, de Gee AJ, Davidson CL. Light initiation of dental resins: dynamics of the polymerization. Biomaterials 1996; 17:23132318. 27. Yoshida K, Greener EH. Effect of photoinitiator on degree of conversion of unfilled light-cured resin. J Dent 1994; 22:296299. 28. Lapcik L, Jancar J, Stasko A, Saha P. Electron paramagnetic resonance study of free-radical kinetics in ultaviolet lightcured dimethacrylate copolymers. J Mater Sci, Mater Med 1998; 9:257262. 29. Sustercic D, Cevc P, Funduk N, Pintar MM. Determination of curing time in visible lightcured composite resins of different thickness by electron paramagnetic resonance. J Mater Sci, Mater Med 1997; 8:507510. 30. Burtscher P. Stability of radicals in cured composite materials. Dent Mater 1993; 9:218221. 31. Ottaviani MF, Fiorini A, Mason PN, Corvaja C. Electron spin-resonance studies of dental composites: effects of irradiation time, decay over time, pulverization, and temperature variations. Dent Mater 1992; 8:118124. 32. Finger WJ, Lee KS, Podszun W. Monomers with low oxygen inhibition as enamel/ dentin adhesives. Dent Mater 1996; 12:256261. 33. Loza-Herrero MA, Rueggeberg FA, Caughman WF, Schuster GS, Lefebvre CA, Gardner FM. Effect of heating delay on conversion and strength of a post-cured resin composite. J Dent Res 1988; 77:426431. 34. Eliades GC, Vougioklakis CG, Caputo AA. Degree of double bond conversion in light-cured composites. Dent Mater 1987; 3:1925. 35. Kildal KK, Ruyter IE. How different curing methods affect the degree of conversion of resin-based inlay/onlay materials. Acta Odontol Scand 1994; 52:315322. 36. Tarumi H, Imazato S, Ehara A, Kato S, Ebi N, Ebisu S. Post-irradiation polymerization of composites containing bis-GMA and TEGDMA. Dent Mater 1999; 15:238242. 37. Shortall AC, Harrington E. Temperature rise during polymerization of light-activated resin composites. J Oral Rehabil 1998; 25:908913.

VOLUME 12, NUMBER 6, 2000

307

CURING DENTAL RESINS AND COMPOSITES BY PHOTOPOLYMERIZATION

38. Vanherle G, Lambrechts P, Bream M. Overview of the clinical requirements for posterior composites. In: Vanherle G, Smith DC, eds. Posterior composite resin restorative materials. Minneapolis: Minnesota Mining and Mfg. Co., 1985. 39. Ferracane JL. Current trends in dental composites. Crit Rev Oral Biol Med 1995; 6:302318. 40. Stansbury JW, Antonucci JM. Dimethacrylate monomers with varied fluorine contents and distributions. Dent Mater 1999; 15:166173. 41. Patel MP, Davy KWM, Braden M. Refractive-index and molar refraction of methacrylate monomers and polymers. Biomaterials 1992; 13:643645. 42. Kawaguchi M, Fukushima T, Miyazaki K. The relationship between cure depth and transmission coefficient of visible lightactivated resin composites. J Dent Res 1994; 73:516521. 43. Nomoto R. Effect of light wavelength on polymerization of light-cured resins. Dent Mater J 1997; 16:6073. 44. Kloosterboer JG, Lijten GFCM, Boots HMJ. Network formation by chain crosslinking photopolymerization and some applications in electronics. Makromol Chem, Macromol Symp 1989; 24:223230. 45. Swartz ML, Phillips RW, Rhodes B. Visible light-activated resins: depth of cure. J Am Dent Assoc 1983; 106:634637.

46. Harrington E, Wilson HJ, Shortall AC. Light-activated restorative materials: a method of determining effective radiation times. J Oral Rehabil 1996; 23:210218. 47. Nie J, Andrzejewska E, Rabek JF, et al. Effect of peroxides and hydroperoxides on the camphorquinone-initiated photopolymerization. Macromol Chem Phys 1999; 200:16921701. 48. Park YJ, Chae KH, Rawls HR. Development of a new photoinitiation system for dental light-cure composite resins. Dent Mater 1999; 15:120127. 49. Peutzfeldt A, Asmussen E. Effect of propanal and diacetyl on quantity of remaining double bonds of chemically cured BisGMA/TEGDMA resins. Eur J Oral Sci 1996; 104:309312. 50. Guggenberger R, Lechner G, Weinmann W. Bisacylphosphine oxides as photoinitiators in dental materials. Polym Prepr (Am Chem Soc) 1997; 38:105106. 51. Padon KS, Scranton AB. The effect of oxygen on the three-component radical photoinitiator system: methylene blue, N-methyldiethanolamine, and diphenyliodonium chloride. J Polym Sci, Polym Chem Ed 2000; 38:33363346. 52. Hayakawa T, Horie K. Effect of watersoluble photoinitiator on the adhesion between composite and tooth substrate. Dent Mater 1992; 8:351353. 53. Hayakawa T, Kikutake K, Nemoto K. Effectiveness of the addition of water-

soluble photoinitiator into the self-etching primers on the adhesion of a resin composite to polished dentin and enamel. Dent Mater J 1999; 18:324333. 54. Imai Y, Kadoma Y, Kojima K, Akimoto T, Ikakura K, Otha T. Importance of polymerization initiator systems and interfacial initiation of polymerization in adhesive bonding of resin to dentin. J Dent Res 1991; 70:10881091. 55. Tilbrook DA, Clarke RL, Howle NE, Braden M. Photocurable epoxy-polyol matrices for use in dental composite I. Biomaterials 2000; 21:17431753. 56. Kannurpatti AR, Anseth JW, Bowman CN. A study of the evolution of mechanical properties and structural heterogeneity of polymer networks formed by photopolymerizations of multifunctional (meth)acrylates. Polymer 1998; 39:25072513.

Reprint requests: Jeffrey W. Stansbury, PhD, University of Colorado Health Sciences Center, Biomaterials Research Center, P.O. Box 6508/Mail stop F436, Aurora, CO 80045-0508; e-mail: jeffrey.stansbury@ uchsc.edu 2000 BC Decker Inc

308

JOURNAL OF ESTHETIC DENTISTRY

Polymerization Contraction Stress of Resin Composite Restorations in a Model Class I Cavity Configuration Using Photoelastic Analysis
YOSHIFUMI KINOMOTO, DDS, PHD* MITSUO TORII, DDS, PHD FUMIO TAKESHIGE, DDS, PHD* SHIGEYUKI EBISU, DDS, PHD*

ABSTRACT

Purpose: An important factor that contributes to deterioration of resin composite restorations is contraction stress that occurs during polymerization. The purpose of this article is to familiarize the clinician with the characteristics of contraction stress by visualizing the stresses associated with this invisible and complex phenomenon. Materials and Methods: Internal residual stresses generated during polymerization of resin composite restorations were determined using micro-photoelastic analysis. Butt-joint preparations simulating Class I restorations (2.0 mm 5.0 mm, 2.0 mm in depth) were prepared in three types of substrates (bovine teeth, posterior composite resin, and transparent composite resin) and were used to examine contraction stress in and around the preparations. Three types of composite materials (a posterior composite, a self-cured transparent composite, and a light-cured transparent composite) were used as the restorative materials. The self-cured composite is an experimental material, and the others are commercial products. After treatment of the preparation walls with a bonding system, the preparations were bulk filled with composite. Specimens for photoelastic analysis were prepared by cutting sections perpendicular to the long axis of the preparation. Fringe patterns for directions and magnitudes of stresses were obtained using transmitted and reflected polarized light with polarizing microscopes. Then, the photoelastic analysis was performed to examine stresses in and around the preparations. Results: When cavity preparations in bovine teeth were filled with light-cured composite, a gap was formed between the dentinal wall and the composite restorative material, resulting in very low stress within the restoration. When cavity preparations in the posterior composite models were filled with either self-cured or light-cured composite, the stress distribution in the two composites was similar, but the magnitude of the stress was greater in the light-cured material. When preparations in the transparent composite models were filled with posterior composite and lightcured transparent composite material, significant stress was generated in the preparation models simulating tooth structure, owing to the contraction of both restorative materials.
CLINICAL SIGNIFICANCE

Polymerization contraction stress is an undesirable and inevitable characteristic of adhesive restorations encountered in clinical dentistry that may compromise restoration success. Clinicians must understand the concept of polymerization contraction stress and realize that the quality of composite resin restorations depends on successful management of these stresses. (J Esthet Dent 12:309319, 2000)
*Department of Restorative Dentistry and Endodontology, Osaka University Graduate School of Dentistry, Osaka, Japan Department of Operative Dentistry and Endodontology, Kagoshima University Dental School, Kagoshima, Japan
VOLUME 12, NUMBER 6, 2000

309

POLYMERIZATION CONTRACTION STRESS OF RESIN COMPOSITE RESTORATIONS IN A MODEL CLASS I CAVITY CONFIGURATION USING PHOTOELASTIC ANALYSIS

esin composite restoratives solidify by means of the chemical process termed polymerization. Polymerization is a repetitive intermolecular reaction whereby resin monomer molecules are converted into a network by covalent bonding to each other along polymer chains.1 Contraction shrinkage occurs when the distance between monomer molecules associated with van der Waals forces is decreased as a result of the formation of covalent bonds during monomer conversion. Moreover, the distance change between molecules in fluid state and solid state induces contraction shrinkage. Thus, shrinkage occurs during polymerization of resin composite restorations as a function of the chemistry of the synthetic resins. Shrinkage is not a problem when the resin composite shrinks on flat dentin surfaces. However, if the shrinkage occurs when the resin composite is confined within a tooth cavity preparation, contraction stress occurs as a result of competition between strength of the bond with tooth structure and the contraction forces.25 In clinical dentistry, contraction stress is one of the problems encountered with adhesive restorations and an important factor in the quality of marginal adaptation in resin composite restorations.6 Poor marginal adaptation of a composite restoration results in microleakage and its possible sequelae: thermal sensitivity, pulpal irritation, and secondary caries.

There have been many studies evaluating the quality of marginal adaptation and associated microleakage.712 However, because the methods for determining contraction stress are complex, relatively few studies have examined polymerization contraction stress of resin composite restoration directly. Studies examining contraction stress have been carried out using strain gauge,13 finite elemental analysis (FEA),1416 tensiometer,17,18 and photoelastic methods.19,20 Photoelastic analysis is a technique for transforming internal stresses produced in materials into visible light patterns that indicate the locations and magnitudes of stresses.21,22 In the dental literature, stresses generated within inlays, crowns, posts, and abutments have been examined using photoelastic analysis.2326 Modifications of traditional photoelastic methods using transparent resin composites and microscopes have been made, to visualize contraction stress within resin composite restorations.19,20 The goal of this article is to familiarize the clinician with the characteristics of contraction stress associated with polymerization shrinkage by visualizing the stresses associated with this invisible and complex phenomenon.
MATERIALS AND METHODS

Extracted bovine incisors were stored in sterile water and frozen at 20C for less than 1 month after

removal of the pulpal soft tissue. After tooth thawing, flat enamel surfaces were prepared by grinding the teeth with wet No. 600-grit SiC paper. Two different types of transparent resin composite materials were used in this study: Palfique clear (Tokuyama Co., Tokuyama, Japan) and an experimental pastepaste self-cured resin composite. Palfique clear is a commercial visible light-activated composite material, which is used for splinting of teeth and glazing of composite restorations. The monomer consists of a mixture of 40% Bis-GMA and 60% by weight of triethylene glycol dimethacrylate (TEGDMA). The inorganic filler is 0.8-m spherical silicon dioxide that constitutes 60% of the weight of the composite. Palfique clear becomes almost transparent after polymerization, because the refraction indices of monomer and filler are adjusted to be equivalent. The volumetric polymerization shrinkage of Palfique clear, measured using a dilatometer with a mercury-filled capillary, was 5.86%, and its bending elastic modulus was 3.3 0.4 GPa.27 The experimental paste-paste self-cured resin composite was prepared using the same filler and monomer as light-cured Palfique clear. This experimental self-cured composite was mixed using the bubbleless mixing technique.28 The bending elastic modulus was 3.1 0.6 GPa. The bending moduli of each curing type (n = 6) were determined 30 minutes after the composite materials were packed

310

JOURNAL OF ESTHETIC DENTISTRY

KINOMOTO ET AL

into a brass mold (30 5 2.5 mm) using a three-point bending test (span length 20 mm).20 The difference in elastic moduli between self-cured and light-cured composite materials was not statistically significant (Students t-test). Calibration of the stress optical coefficient, performed as previously reported,19 was determined to be 1.44 1011 m2/N for Palfique clear and 2.07 1011 m2/N for the experimental self-cured composite. Clearfil PhotoPosterior (Kuraray, Osaka, Japan) was the commercial light-cured posterior composite material used in this study. The volumetric polymerization shrinkage of this material was 1.88%, and its bending elastic modulus was 6.1 0.5 GPa, using methods described above. Cavity Preparations in Models Butt-joint cavities simulating Class I preparations (box-shaped cavity preparations) were prepared in bovine teeth or models constructed from the resin composite materials described above. The dimensions of each cavity preparation were 5.0 mm 2.0 mm and 2.0 mm in depth (Figure 1). This preparation has a configuration factor (C-factor) of 3.8 as defined by Feilzer et al.29 In bovine teeth, standardized preparations were made with a diamond point under water cooling using a cavity preparation device.30 Cavity preparations in the resin composite models were constructed from an

Figure 1. Schematic illustrations of the prepared cavity in the resin composite model and the sectioned specimen for photoelastic analysis.

acrylic model. An impression of an acrylic model (5 mm 2 mm 2 mm) was made using silicon impression material. The final composite model preparation was constructed by filling the impression with the appropriate composite material and light-cured for 240 seconds with a light-curing unit (Quick Light, Morita, Osaka, Japan). No marginal defects were observed between the preparation walls and the restoration when these model preparations were filled with the resin composite test materials. Experimental Design Because analysis of internal stresses can be performed only in transparent materials, transparent resin composite restorations were placed in simulated cavity preparations in bovine teeth or resin composite models (Table 1). In group I, lightcured transparent composite restorations were evaluated in extracted bovine teeth to simulate

clinical conditions closely. In group II, the difference between self-cured (II-S) and light-cured (II-L) transparent composite restorations was evaluated in preparations in posterior composite models. Stress distribution around cavity preparations prepared in light-cured transparent composite models and restored with posterior composite (III-P) or light-cured transparent composite material (III-L) was demonstrated in group III. Fabrication of Test Specimens Cavity preparations in bovine teeth were treated with 40% phosphoric acid solution (K-etchant, Kuraray) for 30 seconds, spray-washed, and dried with air. In preparations to be filled with light-cured composite material, a dual-cured dentin bonding agent (Clearfil Photobond, Kuraray) was mixed according to manufacturers instructions, applied to the preparation, and exposed for 20 seconds with a light-curing unit

VOLUME 12, NUMBER 6, 2000

311

POLYMERIZATION CONTRACTION STRESS OF RESIN COMPOSITE RESTORATIONS IN A MODEL CLASS I CAVITY CONFIGURATION USING PHOTOELASTIC ANALYSIS

TABLE 1. SUMMARY OF EXPERIMENTAL DESIGN.

Number Group Preparation Models Restorative Materials Rationale Underlying the Groups of Trial

I II-S II-L III-P III-T

Bovine tooth Light-cured posterior composite* Light-cured posterior composite Light-cured transparent composite Light-cured transparent composite

Light-cured transparent composite Self-cured transparent composite Light-cured transparent composite Light-cured posterior composite Light-cured transparent composite

To simulate clinical conditions closely To examine stresses in self- and light-cured composites in preparation models

7 7 7

To compare stresses around the preparations between posterior and transparent composites

5 5

*Commercial resin composite for restorative use; only self-cured transparent composite is an experimental material; commercial resin composite for splinting and glazing.

(Quick Light). In preparations to be filled with self-cured transparent composite, a self-cured dentin bonding agent (Clearfil New Bond, Kuraray) was applied. After these surface treatments, the preparations were bulk-filled with the appropriate composite material. The free, unbonded surface of each restoration was covered with a celluloid strip (GC, Tokyo, Japan) and a microscope cover slide glass. For light-cured composites, slight pressure was applied to extrude excess material, followed by immediate light exposure for 80 seconds and subsequent water storage at 37C. For self-cured composite, the restorations were stored in air at 37C immediately after covering with a celluloid strip and cover slide. After 30 minutes, specimens were prepared for photoelastic analysis using a water-cooled diamond saw (Isomet, Buehler, Lake

Bluff, Illinois, USA) to cut 2.0-mmthick sections perpendicular to the long axis of the restoration (see Figure 1). Each section was polished on linen with 0.3-m alumina polishing agent (Alpha Micropolish Alumina No. 2, Buehler). Photoelastic Stress Analysis The internal residual stresses generated during polymerization of the restoration were determined using micro-photoelastic analysis. Isoclinic fringe lines (isoclinics) were obtained using a transmitted polarizing microscope (PHO, Nikon, Tokyo, Japan). Colored isochromatic fringe patterns (isochromatics) were obtained using a reflective polarizing microscope (OPTIPHOT, Nikon), placing a metal mirror under the specimens. These images were immediately recorded on black-and-white and color negative films. To obtain isochromatics

without isoclinics, a patched image was prepared from two images set at 45 degrees relative to each other. This process was necessary because isoclinics emerged as dark lines in the isochromatics, because of the use of a plane polarized light source. Although the crowded color fringe patterns indicated a high stress level at a free boundary condition, analysis of the fringe patterns inside the materials by the following method determined the principal stresses in the specimens. Stress trajectories were drawn from isoclinics, using geometric construction, and lines of equal principal stress difference were constructed from the color prints of isochromatics, using the stress-optic law.21,22 Isoclinics were sketched as the locus of points along which the principal stresses were parallel. Stress trajectories indicate the directions of the

312

JOURNAL OF ESTHETIC DENTISTRY

KINOMOTO ET AL

principal stresses at each point. Isochromatics show the locus of points of constant difference between the maximum and minimum principal stresses. Principal stresses along stress trajectories at any given point in the restoration were determined using a graphic integration method. Normal and shear stresses between the composite and the preparation wall were calculated using the direction and the magnitude of principal stresses in the resin composite restoration approximately 0.1 mm from the wall (Figure 2). Data analysis was accomplished using linear regression. Test of Bond Integrity of Restorations To evaluate the integrity of the bond between transparent composite restoration and the preparation wall, specimen cross-sections were placed in a dye solution (0.5% basic fuchsin) for 1 minute and examined for staining under a stereomicroscope (SMZ-10, Nikon) at 10 times magnification. Presence of the dye along the interface after immersion indicated gap formation and failure of the two materials to bond.
RESULTS

Figure 2. The procedure of photoelastic analysis.

stress change at a deeper part of the preparation. Because characteristic colored stress patterns were not apparent by reflective polarized light observation, the internal stresses generated in this group were not large enough to examine in the present method. Fuchsin staining was observed between the dentinal wall and transparent composite restoration in every specimen prepared in this group. No dye was present between enamel and the composite.

In group II, internal stresses generated in the restorations were compared between self-cured (II-S) and light-cured (II-L) composite. The isoclinics of groups II-S and II-L were similar and are presented in Figure 4. Stress trajectories constructed from isoclinics appeared to have similar patterns for both types of composite cure. Isochromatics and equal principal stress differences are shown in Figure 5. Principal stress differences generated in light-cured composite were greater

Because isoclinics and isochromatics obtained from every specimen were similar in each experimental group, representative samples are presented in the results. Isoclinics and stress trajectories for group I (bovine teeth) are shown in Figure 3. Two isotropic points indicated

Figure 3. Isoclinics (A) and stress trajectories of equivalent magnitude (B) in the bovine tooth preparation (group I). E = enamel, D = dentin. Two isotropic points on either side of the preparation walls are apparent near the enameldentin junctions. (Reprinted from Kinomoto Y, Torii M. Photoelastic analysis of polymerization contraction stresses in resin composite restorations. J Dent 1998, with permission from Elsevier Science.)

VOLUME 12, NUMBER 6, 2000

313

POLYMERIZATION CONTRACTION STRESS OF RESIN COMPOSITE RESTORATIONS IN A MODEL CLASS I CAVITY CONFIGURATION USING PHOTOELASTIC ANALYSIS

Figure 4. Isoclinics and stress trajectories of equivalent magnitude in self-cured (group II-S) and light-cured (group II-L) composites. A, isoclinics in self-cured composite; B, stress trajectories in self-cured composite; C, isoclinics in light-cured composited; and D, stress trajectories in light-cured composite. Solid lines are maximum principal stresses and dotted lines are minimum principal stresses along lines of equivalent stress trajectories. (Reprinted from Kinomoto Y, Torii M, Takeshige F, Ebisu S. Comparison of polymerization contraction stresses between self- and light-cured composites. J Dent 1999; 27:383389, with permission from Elsevier Science.)

than in self-cured material. Normal and shear stresses generated at the interface between the composite restoration and the preparation walls are shown in Figure 6. All data obtained were plotted. Stress patterns generated along the preparation walls were similar for both

types of composites, even though they were cured using two different methods. Simple regression indicated that normal and shear stresses at the lateral wall were a linear function of the distance from the cavosurface margin. By contrast, normal stresses on the preparation

floor were a function of distance from the center of the preparation floor defined by curves constructed using second-order polynomial regression. The largest stresses, which are normal stresses, were found near the internal line angle of the cavity preparation in both types

Figure 5. Isochromatics and lines of equal principal stress difference in self-cured (group II-S) and light-cured (group II-L) composites. A, Isochromatics in self-cured composite; B, lines of equal principal stress difference in self-cured composite; C, isochromatics in light-cured composite; and D, lines of equal principal stress difference in light-cured composite. To obtain isochromatics without isoclinics, a patched image was prepared from two images set at 45 degrees relative to each other. Higher numbers represent greater stress difference values. (Reprinted from Kinomoto Y, Torii M, Takeshige F, Ebisu S. Comparison of polymerization contraction stresses between self- and light-cured composites. J Dent 1999; 27:383389, with permission from Elsevier Science.)

314

JOURNAL OF ESTHETIC DENTISTRY

KINOMOTO ET AL

of composite, reaching 23 MPa in light-cured composite and 12 MPa in self-cured composite. These stresses were defined at a point 0.25 mm from the internal line angle to avoid an edge effect. In group III, stresses generated around the preparations were compared between posterior composite (III-P) and light-cured transparent composite (III-T). Isoclinics and stress trajectories in the preparation models were similar in these two materials (Figure 7). The isochromatic pictures indicate more fringes in the model preparations in group III-T than in group III-P.
DISCUSSION

In this investigation, the distributions and magnitudes of internal stresses generated in and around resin composite restorations resulting from polymerization shrinkage were examined using micro-photoelastic analysis. For simplicity, a rectangular parallelepiped and an axisymmetric preparation were used to analyze internal stress development. It is assumed that the maximum contraction stress occurs between the longitudinal walls of the prepared cavity. The stress profile obtained from analysis of this preparation type should closely approximate that generated between the opposing walls of a Class I restoration. Many assumptions and simplifications underlie this type of stress analysis. A model preparation sys-

Figure 6. Interfacial normal (solid symbols) and shear (hollow symbols) stresses generated between the transparent composite and the preparation wall. Square symbols are data in self-cured composite and circle symbols are data in light-cured composite. In selfcured composite (group II-S): lateral wall (AB)normal stresses (stresses perpendicular to the wall, pulling restoration toward the restoration center) y = 5.740x + 2.124 (r = 0.906); shear stresses (stresses parallel to the wall, pulling restoration upward, toward the unbonded, top surface) y = 3.074x + 6.782 (r = 0.944); preparation floor (BC)normal stresses (stresses pulling restoration up, away from the preparation floor toward the unbonded, top surface) y = 7.456x2 0.216x + 7.107 (r2= 0.919). In light-cured composite (group II-L): lateral wall (AB)normal stresses y = 8.556x + 6.165 (r = 0.918), shear stresses y = 4.766x + 10.796 (r = 0.866); preparation floor (BC)normal stresses y = 17.227x2 + 0.298x + 12.762 (r2 = 0.984). (Reprinted from Kinomoto Y, Torii M, Takeshige F, Ebisu S. Comparison of polymerization contraction stresses between self- and light-cured composites. J Dent 1999; 27:383389, with permission from Elsevier Science.)

tem consisting of posterior composite material was employed to obtain sufficient bonding between the restoration and the preparation walls. Furthermore, because the test composite was transparent after polymerization, neither light attenuation through the material nor the distribution of the degree of cure was considered in the analysis. The elastic modulus of the posterior composite falls between those of human enamel and dentin.3133 Finally, the transparent composites have lower viscosity than restorative composites and may have higher flow capacity in the gel stage.34 Although these factors may

have some influence on the interpretation of the stress analysis, it is unlikely that the distribution of the internal stresses was affected. Therefore, comparison of the internal stresses resulting from polymerization shrinkage in the different cavity preparation conditions may be considered a reflection of the stresses generated under clinical conditions. In group I, model preparations were made in bovine teeth to simulate clinical conditions more closely, particularly interactions of the resin material with enamel and dentin. In this group, the bond between the

VOLUME 12, NUMBER 6, 2000

315

POLYMERIZATION CONTRACTION STRESS OF RESIN COMPOSITE RESTORATIONS IN A MODEL CLASS I CAVITY CONFIGURATION USING PHOTOELASTIC ANALYSIS

dentinal wall and the transparent composite failed, but the bond between enamel and composite was maintained. An isochromatic color fringe, indicating the difference of principal stresses, was not clearly observed in these specimens, probably because the internal stress was reduced, owing to the development of the gap between the dentinal wall and resin composite. This gap

played an important role in stress relief and increased the flow capacity of the composite during polymerization, since the free surface of the restoration was presented not only at the outer surface but also at the gap along the dentinal wall. Because internal stress was reduced in the tooth preparation as a result of gap formation, the composite

model preparation system was used for group II to explore the contraction stress generated when composite polymerized under confined conditions. Comparison between groups II-S and II-L indicated that the stress distributions in the self- and light-cured composites were similar, but the stress magnitudes were different, with the light-cured group demonstrating higher stress values.

Figure 7. Stress trajectories, isochromatics, and lines of equal principal stress difference with the posterior composite (group III-P) and the transparent composite (group III-T) in the transparent preparation models. The rationale underlying this combination was to compare stresses around the preparations between the posterior composite and the transparent composite. A, Stress trajectories in group III-P; B, isochromatics in group III-P; C, lines of equal principal stress difference in group III-P; D, stress trajectories in group III-T; E, isochromatics in group III-T; and F, lines of equal principal stress difference in group IIIT. Solid lines connect equivalent maximum principal stresses and dotted lines connect equivalent minimum principal stresses in stress trajectories. To obtain isochromatics without isoclinics, a patched image was prepared from two images set at 45 degrees relative to each other. Higher numbers represent greater stress difference values.

316

JOURNAL OF ESTHETIC DENTISTRY

KINOMOTO ET AL

It was generally agreed that lightcured composites shrink toward the direction of the light source, whereas shrinkage occurs toward the center of the composite within self-cured composites. Versluis et al, using finite element analysis, reported little difference in shrinkage vector patterns between self- and lightcured composites.16 Results of the present study are consistent with this report, although the data analyzed internal stresses rather than shrinkage vectors. It is concluded that stress distribution is predominantly controlled by boundary conditions (such as free surface, intact bond, and gap formation) and not by the different modes of curing. In group II-L, the light-cured composite, the larger normal stress occurred around the internal line angle, reaching values as high as 23 MPa. Stresses generated along the preparation lateral wall were not uniformly distributed, with the greatest stresses occurring within deeper locations. Consistent with reports by Versluis et al,16 shrinkage vectors were oriented toward the deeper, restricted margin, not toward the light source in the light-cured composite. The magnitude of normal tensile stress obtained in this study was lower than that reported by Rees and Jacobsen and Mahler14,35 and larger than that estimated by Bowen and colleagues and Davidson and De Gee.3,36 These conflicting results may be attributed to the

several assumptions and simplifications implicit in the model system for stress analysis used in this study. They may also be related to the fact that different analytic methods were used to determine stress. Considering the low elastic modulus of the transparent composite, as well as the larger volumetric polymerization shrinkage contributing to reduce internal stresses, the stress generated in the clinical condition may be larger than that determined in the present study. The magnitude of stress difference between self- and light-cured composites has been observed previously.37 The main reason for this magnitude difference is that the polymerization rate of light-cured composite is greater than that of self-cured material. In a clinical situation, it takes only 40 seconds for a 2-mm-thick layer of light-cured composite to cure, whereas selfcured composite takes several minutes. The polymerization rate for these two types of composites is obviously different. Because lightcured composite exists in a gel stage only for a moment, there is not enough time for the resin to flow. Studies focusing on polymerization rate indicate a reduction in the width and extension of the marginal gaps in light-cured composite restorations at reduced rate.38,39 The light exposure method termed stepped light intensity or soft-

start polymerization has been advocated for composite cure.40,41 This method uses an initial lowoutput light intensity followed by a greater intensity. The results of the present study suggest that it may be reasonable to adjust the polymerization rate so that it is slower during the initial step and faster for the remainder of the process. However, the first low-intensity step is usually 10 or 20 seconds in duration.42 Compared to a curing duration of several minutes for a self-cured composite, this time seems to be of short duration. It has not been sufficiently demonstrated that the reduction of polymerization rate achieved by these low-intensity methods contributes significantly to stress reduction. Further research regarding the light-exposure method should be performed to clarify the benefit of this method. Analysis of stress distribution differences between the transparent and posterior composite in group III indicated that both groups were similar, but the magnitude of stress, the color fringe patterns, was different. The opposite lateral walls resist forces generated by the contraction of the restoration, resulting in stress concentration at the internal line angles. The fringe of the transparent restoration (group III-T) is slightly greater than that of the posterior composite restoration (group III-P), indicating greater stress development. This difference is likely the

VOLUME 12, NUMBER 6, 2000

317

POLYMERIZATION CONTRACTION STRESS OF RESIN COMPOSITE RESTORATIONS IN A MODEL CLASS I CAVITY CONFIGURATION USING PHOTOELASTIC ANALYSIS

result of the larger polymerization shrinkage of the transparent composite compared to that of the posterior material. Greater deformation of the preparation model generated the larger fringe patterns. The results indicate that distribution of contraction stress generated within the transparent composite is similar to that generated within the clinical posterior material. Also, the results show that contraction stress results from competition between bond strength of the composite to the preparation walls (tooth structure) and the contraction forces associated with polymerization of the restoration itself.
CONCLUSION

tal transparent resin composite material, and Dr. D.L. Carnes, University of Texas Health Science Center at San Antonio, for reviewing and proofing this manuscript.
REFERENCES

11. Hansen EK. Effect of cavity depth and application technique on marginal adaptation of resins in dentin cavities. J Dent Res 1986; 65:13191321. 12. Tay ER, Gwinnett AJ, Pang KM, et al. Variability in microleakage observed in a total-etch wet-bonding technique under different handling conditions. J Dent Res 1995; 74:11681178. 13. Sakaguchi RL, Sasik CT, Bunczak MA, et al. Strain gauge method for measuring polymerization contraction of composite resins. J Dent 1991; 19:312316. 14. Rees JS, Jacobsen PH. Stress generated by luting resins during cementation of composite and ceramic inlays. J Oral Rehabil 1992; 19:115122. 15. Katona TR, Winkler MM. Stress analysis of a bulk-filling Class V light-cured composite restoration. J Dent Res 1994; 73:14701477. 16. Versluis A, Tantbirojn D, Douglas WH. Do dental composites always shrink toward the light? J Dent Res 1998; 77:14351445. 17. Davidson CL, De Gee AJ. Relaxation of polymerization contraction stresses by flow in dental composites. J Dent Res 1984; 63:146148. 18. Choi KK, Condon JR, Ferracane JL. The effect of adhesive thickness on polymerization contraction stress of composite. J Dent Res 2000; 79:812817. 19. Kinomoto Y, Torii M. Photoelastic analysis of polymerization contraction stresses in resin composite restorations. J Dent 1998; 26:165171. 20. Kinomoto Y, Torii M, Takeshige F, Ebisu S. Comparison of polymerization contraction stresses between self- and light-curing composites. J Dent 1999; 27:383389. 21. Frocht MM. Photoelasticity. London: John Wiley & Sons, 1941:144349. 22. Kuske A, Robertson G. Photoelastic stress analysis. London: John Wiley & Sons, 1974:105389. 23. Craig RG, El-Ebrashi K, LePeak PJ, Peyton FA. Experimental stress analysis of dental restorations. Part 1. Two-dimensional photoelastic stress analysis of inlays. J Prosthet Dent 1967; 17:277291.

1. Anusavice KJ. Phillips science of dental materials. 10th Ed. Philadelphia: WB Saunders, 1996:211235. 2. Asmussen E. Composite restorative resins. Composition versus wall-to-wall polymerization contraction. Acta Odontol Scand 1975; 33:337344. 3. Bowen RL, Nemoto K, Rapson JE. Adhesion bonding of various materials to hard tooth tissues: forces developing in composite materials during hardening. J Am Dent Assoc 1983; 106:475477. 4. Hegdahl T, Gjerdet NR. Contraction stresses of composite resin filling materials. Acta Odontol Scand 1977; 35:191195. 5. Davidson CL, de Gee AJ, Feilzer A. The competition between the compositedentin bond strength and the polymerization contraction stress. J Dent Res 1984; 63:13961399. 6. Carvalho RM, Pereira JC, Yoshiyama M, Pashley DH. A review of polymerization contraction: the influence of stress development versus stress relief. Oper Dent 1996; 21:1724. 7. Kemp-Scholte CM, Davidson CL. Complete marginal seal of Class V resin composite restorations effected by increased flexibility. J Dent Res 1990; 69:12401243. 8. Uno S, Shimokobe H. Contraction stress and marginal adaptation of composite restorations in dentinal cavity. Dent Mater J 1994; 13:1924. 9. Staninec M, Kawakami M. Adhesion and microleakage tests of a new dentin bonding system. Dent Mater 1993; 9:204208. 10. Prati C, Simpson M, Mitchem J, Tao L, Pashley DH. Relationship between bond strength and microleakage measured in the same Class I restorations. Dent Mater 1992; 8:3741.

Contraction stress is an undesirable and inevitable characteristic of adhesive restorations encountered in clinical dentistry that may compromise the success of restorations. Although there have been attempts designed to reduce this stress in composite restorations, few methods have been shown to be effective. Therefore, clinicians must understand concepts underlying polymerization contraction stress and realize that the quality of composite resin restorations depends on the successful management of this stress.
DISCLOSURE AND ACKNOWLEDGMENTS

The authors wish to thank Mr. S. Yuasa, Tokuyama Co., Tokuyama, Japan, for providing the experimen-

318

JOURNAL OF ESTHETIC DENTISTRY

KINOMOTO ET AL

24. Farah JW, Craig RG. Stress analysis of three marginal configuration of full posterior crowns by three-dimensional photoelasticity. J Dent Res 1974; 53:12191225. 25. Burns DA, Krause WR, Douglas HB, Burns DR. Stress distribution surrounding endodontic posts. J Prosthet Dent 1990; 64:412418. 26. Standlee JP, Caputo AA. Load transfer by fixed partial dentures with three abutments. Quintessence Int 1988; 19:403410. 27. Iga M, Takeshige F, Ui T, et al. The relationship between polymerization shrinkage measured by a modified dilatometer and the inorganic filler content of light-cured composites. Dent Mater J 1991; 10:3845. 28. Hosoda H, Yamada T, Inokoshi S, et al. Contrivance of bubble-less mixing technique for paste-paste type composite resins. Jpn J Conserv Dent 1985; 28:12661278. 29. Feilzer AJ, De Gee AJ, Davidson CL. Setting stress in composite resin in relation to configuration of the restoration. J Dent Res 1987; 66:16361639. 30. Iwami Y, Yamamoto Y, Ebisu S. A new electrical method for detecting marginal leakage of in vitro resin restorations. J Dent 2000; 28:241247.

31. Craig RG, Peyton FA, Johnson DW. Compressive properties of enamel, dental cements, and gold. J Dent Res 1961; 40:936945. 32. Willems G, Lambrechts P, Braem M, et al. A classification of dental composites according to their morphological and mechanical characteristics. Dent Mater 1992; 8:310319. 33. Sano H, Ciucchi B, Matthews WG, et al. Tensile properties of mineralized and demineralized human and bovine dentin. J Dent Res 1994; 73:12051211. 34. Dauvillier BS, Feilzer AJ, De Gee AJ, Davidson CL. Visco-elastic parameters of dental restorative materials during setting. J Dent Res 2000; 79:818823. 35. Mahler DB. Letter to the editor. J Dent Res 1990; 69:913. 36. Davidson CL, De Gee AJ. Relaxation of polymerization contraction stresses by flow in dental composites. J Dent Res 1984; 63:146148. 37. Feilzer AJ, De Gee AJ, Davidson CL. Setting stresses in composites for two different curing modes. Dent Mater 1993; 9:25. 38. Kato H. Relationship between the velocity of polymerization and adaptation to dentin cavity wall of light-cured composite. Dent Mater J 1987; 6:3237.

39. Uno S, Asmussen E. Marginal adaptation of a restorative resin polymerized at reduced rate. Scand J Dent Res 1991; 99:440444. 40. Mehl A, Hickel R, Kunzelmann KH. Physical properties and gap formation of light-cured composites with and without soft-start polymerization. J Dent 1997; 25:321330. 41. Koran P, Kurschner R. Effect of sequential versus continuous irradiation of a lightcured resin composite on shrinkage, viscosity, adhesion, and degree of polymerization. Am J Dent 1998; 11:1722. 42. Bouschlicher MB, Rueggeberg FA, Boyer DB. Effect of stepped light intensity on polymerization force and conversion in a photoactivated composite. J Esthet Dent 2000; 12:2332.

Reprint requests: Yoshifumi Kinomoto, DDS, PhD, Department of Restorative Dentistry and Endodontology, Osaka University Graduate School of Dentistry, 1-8 Yamadaoka, Suita, Osaka 565-0871, Japan; e-mail: kinomoto@dent.osaka-u.ac.jp 2000 BC Decker Inc

VOLUME 12, NUMBER 6, 2000

319

Effect of Distance on the Power Density from Two Light Guides


RICHARD B. PRICE, DDS, MS, FDS RCS (Edin)* TORE DRAND, DDS, PHD MARY SEDAROUS, BSc PANTELIS ANDREOU, PHD ROBERT W. LONEY, DMD, MS#

ABSTRACT

Purpose: This study determined the effect of distance on the power density from standard and Turbo light guides (Demetron/Kerr, Danbury, Connecticut). Materials and Methods: Power density was measured from 0 to 10 mm away from the tip of standard 8-mm curved light guides and 13/8-mm Turbo curved light guides. To determine the effect of distance on power density, a polynomial regression line was fitted. The Kolmogorov-Smirnov (K-S) statistic and the Wilcoxon rank sum (WR) tests were used to determine if there was a difference in the rate at which the power density decreased for the standard and Turbo light guides as the distance from the tip increased. Photographs of the light dispersion from each tip were also taken. Results: At 0 mm the mean ( SD) power density from the two standard light guides was 743 6.1 mW/cm2 and from the four Turbo light guides was 1128 22.1 mW/cm2. As the distance from the tip of the light-guide tip increased, the power density decreased, but the rate of decrease was greater from the Turbo light guides than from the standard light guides. At 6 mm the power density from the standard light guides fell to 372 mW/cm2 (50% of the original value) and the power density from the Turbo light guides fell to 263 mW/cm2 (23% of the original value). Both the K-S statistic and the WR sum test indicated that the distribution of light intensities was significantly different from the two light guides (WR p-value .0246, K-S p-value < .0001). The two estimated polynomials intersected at 3.66 mm, and the 95% prediction intervals intersected at about 2.8 and 4.8 mm. Therefore, beyond 5 mm away from the tip of the light guide, the standard light guides gave higher power density readings than the Turbo light guides. Photographs showed that the light dispersed at a wider angle from the Turbo light guides than from the standard light guide.
CLINICAL SIGNIFICANCE

Because the design of the light guide affects light dispersion, manufacturers should report the power density at the tip of the light guide and 6 mm from the tip of the light guide (J Esthet Dent 12:320327, 2000)

*Associate Professor, Department of Dental Clinical Sciences, Faculty of Dentistry, Dalhousie University, Halifax, Nova Scotia, Canada Professor and Head, Department of Oral Technology and Dental Materials Science, Dental School, Malm University, Malm, Sweden Research Assistant, Faculty of Dentistry, Dalhousie University, Halifax, Nova Scotia, Canada Biostatistician, Community Health and Epidemiology, Dalhousie University, Halifax, Nova Scotia, Canada # Professor, Department of Clinical Dental Sciences, Faculty of Dentistry, Dalhousie University, Halifax, Nova Scotia, Canada 320
JOURNAL OF ESTHETIC DENTISTRY

PRICE ET AL

he light energy density received by a resin composite from a curing light can be calculated by multiplying the power density at the surface of the resin composite by the duration of light irradiation. If the resin composite restoration does not receive sufficient energy density this may reduce degree of conversion in the monomer,14 increase its cytotoxicity,5 reduce its ultimate hardness,4,611 decrease its dynamic elastic modulus,12 and increase wear and breakdown at the margins.13 It also may leave uncured bonding resin at the bottom of the restoration, resulting in a weak bond to the tooth.14 Light energy is absorbed as the light passes through air.1,1519 Pires et al reported that 2 mm away from the tip of the light guide the power density had fallen to 78% of the original power density; 6 mm away from the tip the power density had fallen to 47% of the original power density.17 In contrast, Prati et al reported that 2 mm away from the light guide, the power density had fallen to 61% of the original power density; 6 mm away the power density had fallen to 23% of the original power density.19 Although it is recommended to place the light tip as close as possible to the surface of the resin composite, clinically this is often difficult to achieve. For example it has been reported that the distance between the cusp tip and the floor of the interproximal box can exceed 7 mm.20 This distance

adversely affects the power density received by the resin composite at the floor of the proximal box. If the restoration is built up incrementally this results in the first increment, furthest away from the light source, receiving a lower energy density than the top increment, closest to the light guide. Surveys have shown that many curing lights used by dentists do not provide adequate energy output.9,21,22 This has been attributed to several factors, including fluctuations in line voltage, age of the bulb, deterioration of the reflector and filter, light-guide wear or contamination, effect of disinfection procedures on light transmission through the curing-light guide, and malfunction of photoconductive fibers in the light guide.7,15 The power density recorded by hand-held radiometers can be used to predict the curing efficiency of a dental curing light.7,23,24 Regular checks for any deterioration of the light-curing unit are recommended using these radiometers to ensure that the curing light produces a power density greater than 300 mW/cm2. To increase depth of cure and decrease the irradiation time, manufacturers have raised the power output from their light sources. One option uses the Turbo light guide (Demetron/Kerr, Danbury, Connecticut) to boost the power density. The Turbo light guide has an 8-mm diameter distal exit aper-

ture but has a larger (13-mm diameter) proximal aperture where the light enters the light guide. This compares to an 8-mm proximal entrance aperture for a standard light guide with an 8-mm distal exit aperture diameter. The individual fiber optic bundles in the Turbo light guide have a larger diameter at the proximal entrance and have a narrower diameter at the distal exit (Figure 1). This concentration provides more fiber optic bundles per square millimeter at the distal exit aperture of the Turbo light guide compared with the standard light guide of the same size, and this boosts the power density output from the Turbo light guide. The Turbo light guide has been reported to increase the degree of conversion at the bottom of 2 mm of resin composite and may cure resin composite 48% faster than when an 11-mm conventional tip on the same curing light is used.25,26 Other developments include high-intensity plasma arc (PAC) lights. One manufacturer claims that 3 seconds of light exposure with their Apollo 95E PAC light (Dental/Medical Diagnostic Systems, Inc., Westlake Village, California) are equivalent to 45 seconds exposure with a conventional quartz tungsten halogen (QTH) light with a power density of at least 500 mW/cm2.27 However, it has been reported that the depth of polymerization for three composites was less when the Apollo 95E PAC light was used for 3 seconds compared to when a conven-

VOLUME 12, NUMBER 6, 2000

321

EFFECT OF DISTANCE ON THE POWER DENSITY FROM TWO LIGHT GUIDES

the power density (mW/cm2) from standard and Turbo light guides. The hypothesis was that as the distance increased, the Turbo light guides would always deliver a greater power density than the standard light guide.
MATERIALS AND METHODS

Figure 1. The entrance and exit apertures showing 1.6 times and 6.4 times views of the fiber optic filaments in the 13/8-mm Turbo light guide. Left, entrance aperture; right, exit aperture.

tional QTH light source was used for 40 seconds.28 Increased microleakage along the dentin margins has also been reported when using a PAC light compared with using a QTH light and 20 seconds irradiation with one PAC light did not polymerize composite as well as when conventional QTH lights were used for 40 seconds.29,30 The power density is usually greatest at the tip of the light guide, but the design of the light guide may alter the relation between power density and distance from the lightguide tip. However, it has been reported that a light-guide tip with a convex profile focused the light output. The power density from this light guide reached a maximum

value at 1 mm from the tip of the light guide and did not fall below the zero distance power density until 3 mm away from the tip of the light guide.16 Additionally, the design of some light guides made with random fiber optic filaments has been shown to produce more even light output across the surface of the light guide compared with light guides manufactured using graded fibers.31 This design difference may also affect the relation between distance and power density. Since the design of the Turbo light guide is different from a standard light guide, the purpose of this study was to investigate whether the Turbo light guide had a focusing effect on the light output and to determine the effect of distance on

One Optilux 500 curing light (serial #5808041 Demetron/Kerr) with an 80 W OptiBulb was used to examine the power density from two standard 8-mm curved light guides (model #20941 Demetron/Kerr) and two 13/8-mm Turbo+ curved light guides (model #952213 Demetron/Kerr). After a pilot project using two Turbo light guides showed a marked decrease in power density as the distance increased, one more 13/8-mm Turbo+ curved light guide (model #952213) and a previous model of 13/8-mm Turbo curved light guide (model #21020 Demetron/Kerr) were also included in the study. The power density outputs from the light guides were all measured using a Cure Rite digital radiometer (Dentsply/Caulk, Milford, Delaware) with a new battery. The Cure Rite radiometer has been shown to be within 2% of a laboratory grade power meter when measuring the power density from a 7.5-mm diameter light guide.32 The Cure Rite has been used by other researchers,17,33 and it provides a digital output in 1 mW/cm2 increments in place of the analog meter found on some radiometers. The power density (mW/cm2) was first measured at

322

JOURNAL OF ESTHETIC DENTISTRY

PRICE ET AL

0 mm with the light guide gently touching the radiometer, and then at 0.5-mm increments in air up to 10 mm away from the radiometer. The curing light was turned on for a period of 20 seconds before taking the radiometer reading. The curing light was allowed to cool between measurements until the fan in the curing light had stopped. The radiometer and the curing light were clamped in a jig that ensured that the tip of the light guide was always parallel to and directly over the radiometer aperture as the light guide was moved away from the radiometer. Therefore, the light guides and the distance between the radiometer and the tip of the light guide were the only variables. The entire sequence of measurements over the 10-mm distance was repeated five times starting from 0 mm for each light guide. The mean power-density readings from the standard and Turbo light guides were plotted to show where the power density outputs from the light guides intersected. To determine the functional form of the dependence of intensity on distance, a polynomial regression line was fitted. For both light guides the Kolmogorov-Smirnov (K-S) statistic and the Wilcoxon rank sum tests were used to determine whether there was a difference in the rate at which the power density decreased as the distance increased (SAS/STAT Software V8.0, SAS Institute Inc. Cary, North Carolina).

To obtain a visual impression of how the light dispersed from the standard and Turbo light guides, photographs were taken of the light dispersing across a piece of black paper placed midway across the diameter of the tip of the light guides. A millimeter scale was included in the photographs, which were taken under standardized lighting conditions and magnification. To determine over what distance it would be clinically relevant to measure power density, 10 Class II preparations were cut by one experienced dentist in extracted human molar teeth of a similar size and cusp form. The preparations were similar in size and shape and were

representative of a typical deep Class II preparation. The gingival box extended just beyond the cementoenamel junction, and there were no pulpal exposures on the pulpal floor. The tip of an 8-mm light guide was positioned over the cusp tip and the distances from the tip of the light guide to the pulpal and gingival floors of the preparations were measured using calipers. The mean ( SD) distance from the cusp tip to the pulpal floor was 4.6 0.5 mm and to the gingival floor was 6.3 0.7 mm (Figure 2).
RESULTS

At 0 mm from the tip of the light guide the mean ( SD) power densities from five measurements from

Figure 2. Mean distances from cusp tip to pulpal and gingival floors.

VOLUME 12, NUMBER 6, 2000

323

EFFECT OF DISTANCE ON THE POWER DENSITY FROM TWO LIGHT GUIDES

two standard light guides were 747 5.1 and 738 3.3 mW/cm2, respectively. From the one Turbo light guide, the power density was 1107 8.6 mW/cm2, and from the three Turbo+ light guides, the power densities were 1159 2.6, 1110 1.9, and 1136 2.3 mW/cm2, respectively. Although the Turbo+ light guides gave slightly greater mean power densities, there was no significant difference between power densities from the Turbo and the Turbo+ light guides (K-S p-value = .3243; Kruskal-Wallis, .5427). Therefore, the data from the Turbo and Turbo+ light guides were combined, and the mean power density from the two standard light guides was 743 6.1 mW/cm2 and that of the four Turbo light guides was 1128 22.1 mW/cm2. The effect of the distance from the light-guide tip on the power density is shown in Table 1, which also shows the percentage decrease from the original power density. The inverse relation between power density and distance is shown in Figure 3. The power density decreased as the distance increased, but for all four Turbo light guides its rate of decline was greater than that from the standard light guides. For the standard light guide, the relation between power density and distance was a third degree polynomial equation with an adjusted R2 = 0.995. For the Turbo light guide, the relation between

power density and distance was a second degree polynomial equation with an adjusted R2 = 0.999. Both the K-S statistic and the Wilcoxon rank sum test showed that the distribution of the power densities over the distance interval (0 to 10 mm away from the tip of the light guide) was significantly different for the standard and Turbo light guides (Wilcoxon p-value = .0246; K-S p-value < .0001). The two estimated polynomials intersected at 3.66 mm from the tip of the light guide. The 95% prediction intervals for the two polynomials intersected at approximately 2.8 and 4.8 mm from the tip of the light guide (see Figure 3). Within this distance interval (2.84.8 mm) the sta-

tistical power to determine whether an observation from the Turbo group would be greater than from the standard group declined from 95% to 5%. Therefore, although initially all the Turbo light guides gave greater power-density readings than the standard light guides, at distances greater than 5 mm away from the tip, the standard light guide delivered a greater power density than the Turbo light guide. The images of the light exiting from the tips of the light guides showed that as the distance from the tip increased the light diffused at a wider angle of about 54 degrees from the Turbo light guides than about 25 degrees from the standard light guide (Figure 4).

TABLE 1. EFFECT OF DISTANCE FROM THE STANDARD AND TURBO LIGHT GUIDES TO THE RADIOMETER ON POWER DENSITY.

Mean Power Density Standard Tip Distance (mm) mW/cm2 SD Percentage of Original Power Turbo Tip mW/cm2 SD % Original Power

0 1 2 3 4 5 6 7 8 9 10

743 6.1 681 6.2 633 6.0 573 3.0 489 3.3 435 3.0 372 4.1 323 2.4 272 3.1 225 5.2 185 1.7

100 92 85 77 66 59 50 44 37 30 25

1128 22.1 944 35.2 782 24.6 613 24.7 470 29.2 351 22.3 263 25.0 194 17.5 142 16.2 101 15.6 68 10.1

100 84 69 54 42 31 23 17 13 9 6

324

JOURNAL OF ESTHETIC DENTISTRY

PRICE ET AL

DISCUSSION

The present experiment supports a previous study that reported that the design of the light guide may alter the relation between power density and distance from the light guide tip.16 To ensure that the results were not attributable to a faulty Turbo or standard light guide, the experiment was repeated using two standard 8-mm curved light guides, one Turbo curved light guide, and three Turbo+ curved light guides. The power density outputs from three Turbo+ light guides and the one Turbo light guide all showed that as the distance increased the radiometer did not always receive more light energy when the Turbo light guide was used, compared with the standard light guide. Although initially the radiometer received more light energy when the Turbo light guide was used, this light energy diffused at a much greater rate than from the standard light guide (see Figures 3 and 4). At a distance of between 2.8 and 4.8 mm away in air from the light tip, the statistical power to determine if the Turbo light guide produced a greater power density than the standard light guide fell from 95% to 5%. Therefore, at distances greater than 4.8 mm, the power density from the standard light guide was greater than that from the Turbo light guide. Therefore, the hypothesis that as the distance increased the Turbo light guides would always deliver more

Figure 3. Effect of distance on power density for the standard and Turbo light guides.

light energy than the standard light guide was rejected. When more intense light energy is used to cure resin composites, more photons hit the camphorquinone photoinitiator molecules within the resin and more photoinitiator molecules are activated and raised to

the triplet or excited state. In this excited state, camphorquinone collides with an amine, and a free radical is formed, which then reacts with the carbon-to-carbon double bond of a monomer molecule and initiates polymerization.34 If insufficient camphorquinone molecules

Figure 4. Light diffusion from a Turbo+, Turbo, and standard light guides.

VOLUME 12, NUMBER 6, 2000

325

EFFECT OF DISTANCE ON THE POWER DENSITY FROM TWO LIGHT GUIDES

are raised to the triplet or activated state, the resin composite is not adequately polymerized. Thus, the power density from the curing light affects the extent and rate of the polymerization reaction.34 The power densities from the three light guides were well above those found from curing lights in many dental offices,9 and they were all more than the minimum 300 to 400 mW/cm2 power density values previously reported as necessary to polymerize resin composite.3,35 Using the same Optilux 500 light source, the 8-mm Turbo light guide boosted the power density output by 52% (1128 22.1 mW/cm2) compared with the 743 6.1 mW/cm2 power density from the 8-mm standard light guide. This result complemented a previous study that reported that using the Turbo light guide increased the degree of conversion of resin composite, but the tip of the light guide was close to the resin composite.25 Different results might have been obtained if the tip had been placed 6 mm away from the resin composite, to simulate a typical clinical distance from the cusp tip to the floor of a proximal box. Using a SureCure radiometer (Ho Dental Co, Goltea, California), it has been reported that 6 mm away from the tip the power density falls to 23% of the original power density.19 This was a greater

reduction than that found in the present study. However, another study, which used a Cure Rite radiometer and an Optilux 401 curing light with an 11-mm light guide, reported that 6 mm away from the tip the power density was 47% of the original power density.17 This was similar to the present study, in which at 6 mm there was a 50% reduction in power density using the 8-mm standard light guide (see Table 1). A 6-mm space may often be encountered clinically between the surface of the resin composite or a light-activated bleaching material and the tip of the light guide. The mean distance of 6.3 0.7 mm from the cusp tip to the gingival floor of a proximal box was similar to the 7-mm distance previously reported.20 This 6-mm space causes a 50% reduction in power density from the standard light guide and a 77% reduction from the Turbo light guide. This may adversely affect the degree of conversion and physical properties of the resin composite or the effectiveness of the bleaching material, especially if the initial power density from the curing light is low (e.g., 400 mW/cm2 will fall to 200 mW/cm2 at 6 mm away from the tip of the standard light guide). This means that there may be many clinical situations in which use of the standard light guide rather than the Turbo light guide would be advisable.

CONCLUSIONS

1. At 0 mm from the tip of the light guide, the 13/8-mm Turbo light guide increased the power density output by 52% compared with the 8-mm standard light guide. 2. Beyond 5 mm from the tip of the light guide, an 8-mm standard light guide delivered a greater power density than the 13/8-mm Turbo light guide. 3. At 10 mm from the tip of the light guide, the 8-mm standard light guide delivered 37% more light energy than the 13/8-mm Turbo light guide.
DISCLOSURE AND ACKNOWLEDGMENTS

The curing light and light guides were kindly donated by the Kerr Corporation, Danbury, Connecticut, USA. The study was financially supported by the Dalhousie University Alumni Oral Health Research Fund.
REFERENCES

1. Nomoto R, Uchida K, Hirasawa T. Effect of light intensity on polymerization of light-cured composite resins. Dent Mater J 1994; 13:198205. 2. Rueggeberg FA, Caughman WF, Curtis JW Jr, Davis HC. Factors affecting cure at depths within light-activated resin composites. Am J Dent 1993; 6:9195. 3. Rueggeberg FA, Caughman WF, Curtis JW Jr. Effect of light intensity and exposure duration on cure of resin composite. Oper Dent 1994; 19:2632. 4. Eliades GC, Vougiouklakis GJ, Caputo AA. Degree of double-bond conversion in light-cured composites. Dent Mater 1987; 3:1925.

326

JOURNAL OF ESTHETIC DENTISTRY

PRICE ET AL

5. Caughman WF, Caughman GB, Shiflett RA, Rueggeberg F, Schuster GS. Correlation of cytotoxicity, filler loading, and curing time of dental composites. Biomaterials 1991; 12:737740. 6. Atmadja G, Bryant RW. Some factors influencing the depth of cure of visible light-activated composite resins. Aust Dent J 1990; 35:213218. 7. Shortall AC, Harrington E, Wilson HJ. Light-curing unit effectiveness assessed by dental radiometers. J Dent 1995; 23:227232. 8. Baharav H, Abraham D, Cardash HS, Helft M. Effect of exposure time on the depth of polymerization of a visible lightcured composite resin. J Oral Rehabil 1988; 15:167172. 9. Pilo R, Oelgiesser D, Cardash HS. A survey of output intensity and potential for depth of cure among light-curing units in clinical use. J Dent 1999; 27:235241. 10. Davidson-Kaban SS, Davidson CL, Feilzer AJ, de Gee AJ, Erdilek N. The effect of curing-light variations on bulk curing and wall-to-wall quality of two types and various shades of resin composites. Dent Mater 1997; 13:344352. 11. Correr Sobrinho L, De Goes MF, Consani MA, Sinhoreti MAC, Knowles JC. Correlation between light intensity and exposure time on the hardness of composite resin. J Mat Sci Mat Med 2000; 11:361364. 12. Harris JS, Jacobsen PH, ODoherty DM. The effect of curing light intensity and test temperature on the dynamic mechanical properties of two polymer composites. J Oral Rehabil 1999; 26:635639. 13. Ferracane JL, Mitchem JC, Condon JR, Todd R. Wear and marginal breakdown of composites with various degrees of cure. J Dent Res 1997; 76:15081516. 14. Lee SY, Greener EH. Effect of excitation energy on dentine bond strength and composite properties. J Dent 1994; 22:175181.

15. Sakaguchi RL, Douglas WH, Peters MC. Curing light performance and polymerization of composite restorative materials. J Dent 1992; 20:183188. 16. Shortall AC, Harrington E. Effectiveness of battery powered light activation units. Br Dent J 1997; 183:95100. 17. Pires JA, Cvitko E, Denehy GE, Swift EJ Jr. Effects of curing tip distance on light intensity and composite resin microhardness. Quintessence Int 1993; 24:517521. 18. Rueggeberg FA, Jordan DM. Effect of light-tip distance on polymerization of resin composite. Int J Prosthodont 1993; 6:364370. 19. Prati C, Chersoni S, Montebugnoli L, Montanari G. Effect of air, dentin and resin-based composite thickness on light intensity reduction. Am J Dent 1999; 12:231234. 20. Yearn JA. Factors affecting cure of visible light activated composites. Int Dent J 1985; 35:218225. 21. Martin FE. A survey of the efficiency of visible light curing units. J Dent 1998; 26:239243. 22. Poulos JG, Styner DL. Curing lights: changes in intensity output with use over time. Gen Dent 1997; 45:7073. 23. Peutzfeldt A. Correlation between recordings obtained with a light-intensity tester and degree of conversion of a light-curing resin. Scand J Dent Res 1994; 102:7375. 24. Rueggeberg FA. Precision of hand-held dental radiometers. Quintessence Int 1993; 24:391396. 25. Curtis JW Jr, Rueggeberg FA, Lee AJ. Curing efficiency of the Turbo Tip. Gen Dent 1995; 43:428433. 26. Clinical Research Associates. Resin curing lights: rapid cure. Clinical Research Associates Newsletter 2000; 24:14. 27. Apollo 95E 1-3 Second Curing Instructions. Westlake Village, CA: Dental/ Medical Diagnostic Systems, 1998:117.

28. Peutzfeldt A, Sahafi A, Asmussen E. Characterization of resin composites polymerized with plasma arc curing units. Dent Mater 2000; 16:330336. 29. Brackett WW, Haisch LD, Covey DA. Effect of plasma arc curing on the microleakage of Class V resin-based composite restorations. Am J Dent 2000; 13:121122. 30. Burgess JO, DeGoes M, Walker R, Ripps AH. An evaluation of four light-curing units comparing soft and hard curing. Pract Periodont Aesthet Dent 1999; 11:125132. 31. Harrington L, Wilson HJ. Determination of radiation energy emitted by light activation units. J Oral Rehabil 1995; 22: 377385. 32. Leonard DL, Charlton DG, Hilton TJ. Effect of curing-tip diameter on the accuracy of dental radiometers. Oper Dent 1999; 24:3137. 33. Versluis A, Tantbirojn D, Douglas WH. Do dental composites always shrink toward the light? J Dent Res 1998; 77:14351445. 34. Rueggeberg F. Contemporary issues in photocuring. Compendium 1999; 20:S415. 35. Curing radiometer operating instructions. Danbury, CT: Demetron Research Corp., 1990.

Presented at the Canadian Academy of Restorative Dentistry and Prosthodontics Annual Scientific Session, September 1416, 2000, Halifax, Nova Scotia, Canada. Reprint requests: Dr. Richard Price, DDS, MS, FDS RCS (Edin), Associate Professor, Department of Dental Clinical Sciences, Faculty of Dentistry, Dalhousie University, Halifax, Nova Scotia, Canada, B3H 3J5; e-mail: rbprice@is.dal.ca 2000 BC Decker Inc

VOLUME 12, NUMBER 6, 2000

327

Effect of Ramped Light Intensity on Polymerization Force and Conversion in a Photoactivated Composite
MURRAY R. BOUSCHLICHER, DDS, MS* FREDERICK A. RUEGGEBERG, DDS, MS

ABSTRACT

Purpose: This study evaluated the effect of ramped light intensity on the polymerization shrinkage forces and degrees of conversion (DC) of a hybrid composite. Materials and Methods: Composite samples were bonded between two steel rods (2.50 mm diameter, 1.25 mm apart, configuration factor = 1.0) mounted in a universal testing machine using a constant displacement mode. Polymerization contraction force was recorded for 250 seconds under four light exposure conditions: group 1, STD: (40 s 800 mW/cm2); group 2, EXP: (150 mW/cm2 logarithmic increase to 800 mW/cm2 over 15 s) + (25 s 800 mW/cm2); group 3, 2-STEP: (10 s 150 mW/cm2) + (30 s 800 mW/cm2); group 4, MED: (80 s 400 mW/cm2). Maximum curing force (N250s) and maximum force rate of the four groups were compared using one-way analysis of variance (ANOVA) ( = 0.05) and the Tukey test. Degrees of conversion obtained with STD, EXP, and MED cure modes were evaluated at three depths (top surface, 1 mm, and 2 mm) using Fourier transform infrared spectroscopy (FTIR). Results: Maximum rates of polymerization shrinkage force development and standard deviations (SD) in ascending order were group 4, MED: 0.33 0.03 N/s; group 2, EXP: 0.35 0.06 N/s; group 1, STD: 0.44 0.03 N/s; and group 3, 2-STEP: 0.46 0.07 N/s. Maximum rates of polymerization shrinkage force development of group 2, EXP and group 4, MED were statistically equivalent and lower than group 1, STD and group 3, 2-STEP. Maximum shrinkage forces ( SD) in ascending order were group 2, EXP: 20.4 2.5 N; group 4, MED: 25.8 1.0 N; group 3, 2-STEP: 27.4 5.8 N, and group 1, STD: 30.5 2.7 N. Maximum force of the EXP mode was statistically lower than MED, 2-STEP, and STD curing modes. The EXP ramp was successful in reducing the conversion rate at the top surface and at 1.0-mm depth, but it did not affect the total conversion compared to the STD 40-second cure mode. There was no difference in DC at the top surface and 1-mm depth with mode of cure. The MED cure mode resulted in a higher DC than the EXP mode at a depth of 2 mm.
CLINICAL SIGNIFICANCE

Maximum shrinkage force and force rate exhibited during the first 250 seconds of polymerization were significantly lower using a ramped light intensity exposure. Ramped light intensity decreased conversion rate at the top surface and at 1.0-mm depth, and did not affect the total extent of conversion compared to a standard 40-second, single-intensity cure mode. The slower conversion rate resulting from ramped light intensity helped to reduce the rate and maximum polymerization stress, but would not be expected to compromise the physical properties for the restorative material, since similar degrees of conversion were obtained. (J Esthet Dent 12:328339, 2000)
*Assistant Professor, Department of Operative Dentistry, College of Dentistry, The University of Iowa, Iowa City, Iowa Professor, Section Director, Dental Materials, Department of Oral Rehabilitation, School of Dentistry, Medical College of Georgia, Augusta, Georgia 328
JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

anufacturers have introduced conventional quartz tungsten halogen (QTH), xenon plasma arc (PAC), and argon ion laser curinglights, with increasingly higher light intensities or power densities (PD, units = mW/cm2) for photoinitiation of resin composite polymerization. Use of a higher PD for fixed exposure durations generally results in a higher degree of monomer conversion to polymer with associated improvements in the mechanical properties of resin composite being polymerized.1 However, higher polymerization shrinkage force rates generated by more rapid conversion may result in loss of adhesion at the toothrestoration interface. If contraction forces exceed the bond strength at this interface, the resulting interfacial gap can lead to staining, marginal leakage, tooth sensitivity, and recurrent caries.24 If this interface remains intact, contraction forces may transfer stress to surrounding tooth structure.5,6 If the surrounding tooth structure is compliant, the resulting strain or deformation may cancel a portion or all of the developing stress. However, in high configurationfactor (C-factor) restorations, fractures in surrounding high-modulus enamel have been reported.7 Whereas an elevated PD is desirable for achieving high conversion and improved mechanical properties in the shortest possible exposure interval, lower light intensities over longer exposure intervals are more

favorable for maintenance of the toothrestoration interface.8 Lower light intensity minimizes marginal gap formation,9 with less residual stress transfer.10 This finding presumably results from a more extended period of viscous flow in the pre-gelation phase within the setting resin similar to that observed in chemical-cure composites. Within limits, it is possible to compensate for a lower PD by increasing exposure duration to obtain an energy density (ED, units = mJ/cm2 = mW/cm2 s) that results in an equivalent degree of conversion (DC).9 However, the increased exposure duration required to obtain an ED resulting in an equivalent DC may be objectionable to clinicians from a time utilization standpoint. When an initial, low PD is increased to a higher, final PD by either a stepped or ramped transition, the resulting soft start energy application sequence may result in improved marginal adaptation and a DC equivalent to that of a single, continuous-intensity energy application sequence. Several studies report improved marginal adaptation with stepped or ramped intensity increases in PD that may improve marginal adaptation by allowing viscous flow of the material during polymerization.8,1114 Commercially available light-curing units with EAS that transition from a low to high PD include a stepped increase (Elipar Highlight, ESPE America, Norristown, Pennsylvania), a dis-

continuous or interrupted step increase (VIP or Variable Intensity Polymerizer, Bisco, Inc., Schaumburg, Illinois), and a logarithmic increase (Elipar Trilight, ESPE America). The purpose of this study was to evaluate the stress development (maximal stress value and stress rate) as well as conversion rate of a commercially available, photoinitiated resin composite when exposed to a variety of curing-light exposure profiles. It was hypothesized that light exposure methods applying initial low power density values would provide significantly lower stress values than a continuous, high-intensity application.
MATERIALS AND METHODS

Threaded steel rods were machined to a cylindrical cross-section 2.50 mm diameter and 15.0 mm height. The flat bonding surfaces were microetched with 30 silanized silica (CoJet-Sand, ESPE America), silanized with ESPE-Sil (ESPE America), and allowed to air dry for 60 seconds. The rods were attached to the load cell and crosshead of a universal testing machine (UTM, MTS 318 Biaxial Test System, MTS Systems Corp., Eden Prairie, Minnesota) in a tensile test configuration with an extensometer maintaining constant displacement of 1.250 0.001 mm (Figure 1). An unfilled bonding resin (All Bond 2 D/E Resin, Bisco, Inc.) was applied to the silanated rod ends, brush thinned, and light-

VOLUME 12, NUMBER 6, 2000

329

BOUSCHLICHER AND RUEGGEBERG EFFECT OF RAMPED LIGHT INTENSITY ON POLYMERIZATION FORCE AND CONVERSION IN A PHOTOACTIVATED COMPOSITE

Figure 1. MTS Model 318 Biaxial Test System.

cured for 20 seconds with a curing light (Elipar Trilight, ESPE America) using a standardized output (800 mW/cm2). Light intensity output was measured with a visible light radiometer (Nova Laser Power/ Energy Monitor, Ophir Optronics Jerusalem Ltd., Jerusalem, Israel). The ends of the vertically oriented rods opposed one another, their parallel end surfaces defining two 2.50-mm diameter (d) disks. Polymerization force was evaluated at a C-factor of 1.0. Configuration factor, the ratio of bonded to unbonded surface areas, was measured according to the method of Feilzer et al.15 When the upper and lower rods approximated each other, the instrument was zeroed. This measurement reference allowed accurate positioning of the rod ends at the desired distance or

height (h = 1.25 mm, d = 2.50 mm, C-factor [C] = d/2 h = 1.0). All tests were conducted at room temperature (23.5 1C). Following surface treatment, the two ends were positioned at a distance greater than h allowing access for loading a hybrid composite (Pertac II, shade A2, ESPE America, Norristown, Pennsylvania) into a movable matrix positioned on the lower rod. Composite specimens were placed under illumination of overhead fluorescent light, as the amount of premature composite curing attributable to ambient fluorescent lighting has been shown to be negligible.16 Following composite placement, the inter-rod distance was reset to h = 1.25 mm, establishing the pre-set C-factor of 1.0. The matrix was then rapidly slid up

onto the upper rod, which removed any excess composite and formed a cylindrically shaped specimen of uncured composite. The extensometer was attached to the upper and lower rods, and the light guide was positioned at a 1.0-mm tip-tosample distance (see Figure 1). The UTM was switched to extensometer displacement control, and light polymerization was initiated. Polymerization contraction force (N), elapsed time, and extensometer displacement were recorded at 0.1-second intervals over a period of 250 seconds from initiation of curing exposure. Light unit onset, intensity change, and light cessation as monitored by a photocell voltage change were recorded on a separate data channel. The means of the maximum force rate development (N/s) and maximum contraction force (N250s) were calculated from the force:time curves of all trials for each experimental group (n = 5/group). Samples were polymerized using four different cure modes, or energy application sequences (EASs). Two single-intensity curing modes (Elipar Trilight) were tested: a conventional high-power density curing mode (STD, 800 mW/cm2) with a 40second exposure duration, and an intermediate-power density (MED, 400 mW/cm2) with an 80-second exposure duration. These two singleintensity curing modes had equivalent energy densities (800 mW/cm2 40 s) = (400 mW/cm2 80 s) = 32.0 Joules/cm2. The two variable-

330

JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

intensity EASs included a steppedPD EAS (STEP, Elipar Highlight, ESPE America) [100 mW/cm2 10 s] + [800 mW/cm2 30 s] and a ramped-PD EAS (EXP, Elipar Trilight) [150 mW/cm2 with a logarithmic increase to 800 mW/cm2 in 15 s] + [800 mW/cm2 25 s]. Power density and total energy per EAS (units = Joules [J]) was measured with a visible-light radiometer (Nova Laser Power/Energy Monitor). The polymerization force rates (maximum slope, N/s) and force maxima (N250s) for each of the four EAS groups (n = 5, total of 20 specimens) were compared with a one-way ANOVA. Tukeys HSD post hoc test was used for pair-wise comparison, and the mean values were grouped into homogeneous subsets. All statistical testing was performed at a pre-set alpha of 0.05. Monomer Conversion Maximal rate (DC/s) and degree of conversion during the first 100 seconds from onset of light-curing exposure (DC100s) (three curing modes or EASs) were determined at a depth of 0.0 mm (top surface), 1.0 mm, and 2.0 mm using Fourier transform infrared spectroscopy (FTIR) with a KRS-5 total attenuated reflectance crystal. Real-time cure kinetics was obtained from the composite at a depth of 0.0 (top surface), 1.0, and 2.0 mm, using newly developed methodology.17 This methodology provides conversion values every 1.7 seconds, as

well as the conversion rate (DC/s). The maximal cure value (DC100s) following light-curing also was obtained. Infrared energy entered the crystal on one beveled surface, reflected off the surface at specific points, and then exited (Figure 2). As a result of differences in the refractive indices between the crystal and the resin, the infrared beam penetrated the composite applied to the surface for only a few microns. The infrared beam exited the crystal after being attenuated by contact with the resin material. The resulting spectrum was then analyzed for characteristics of the resin prior to curing, and then again after the polymerization process was activated. The resin spectrum was obtained for a depth of approximately 5 microns against the crystal face. The infrared spec-

trum of the specimen was obtained in a real-time format during the light exposure. The spectrometer collected four spectra at a resolution of 4 cm1 and averaged them to provide a single spectrum during the time period of 1.7 seconds. This process was continued throughout the entire exposure for a total of approximately 100 seconds. Thus, a set of 40 such totaled and averaged scans was obtained that represents conversion values at specific intervals during and immediately following light-activation. The extent of cure (monomer conversion into polymer) was determined using methods previously published.18 This methodology compared changes in the ratio between the aromatic carbon-to-carbon (C=C) double bond and that of the aliphatic C=C before and during

Figure 2. KRS-5 total attenuated reflectance crystal and resin.

VOLUME 12, NUMBER 6, 2000

331

EFFECT OF RAMPED LIGHT INTENSITY ON POLYMERIZATION FORCE AND CONVERSION IN A PHOTOACTIVATED COMPOSITE

curing. Three replications for each test condition were performed. Statistical comparisons involved the total conversion (%) at the end of the light-curing cycle as well as the maximum cure rate (%/s) during the first few seconds of light activation. A two-way ANOVA was performed to determine the effect of cure mode and depth on conversion 100 seconds after light initiation. A two-way ANOVA also was performed to determine the effect of cure mode and depth on peak conversion rate. If interactions were present between cure mode and depth, oneway ANOVAs were performed to examine the differences. The TukeyKramer post hoc test was applied for pair-wise group comparisons. All statistical testing was performed at a pre-set alpha of 0.05.
RESULTS

Force Rate and Maximum Force Force:time curves were S-shaped (Figure 3). Specimens exposed to STEP and EXP modes exhibited longer delays before contractive force was recorded. Except for the 10-second shift of the STEP

force:time curve on the time axis, which roughly corresponds to the low light-intensity portion of the EAS, the initial 40 seconds of STD and STEP curves had similar slopes with maximum force rates of 0.44 0.03 N/s and 0.46 0.07 N/s, respectively. Maximum force rate development ( SD) in ascending order were group 4, MED: 0.33 0.03 N/s; group 2, EXP: 0.35 0.06 N/s; group 1, STD: 0.44 0.03 N/s; and group 3, 2-STEP: 0.46 0.07 N/s. The maximum rates of polymerization shrinkage force development of group 2, EXP and group 4, MED were statistically equivalent and lower than group 1, STD and group 3, 2-STEP (Table 1). The maximum rates of STD and 2-STEP curing modes were statistically similar. Curing with the EXP mode decreased the maximum force rate development to 80% of the STD rate. Curing with an intermediatepower density (400 mW/cm2) and an 80-second exposure (providing equivalent energy density) reduced the force rate development to 75% of the STD groups maximum rate.

The maximum rate of the 2-STEP mode was statistically equivalent and 104% of the STD rate. Maximum shrinkage forces ( SD) in ascending order were group 2, EXP: 20.4 2.5 N; group 4, MED: 25.8 1.0 N; group 3, 2-STEP: 27.4 5.8 N; and group 1, STD: 30.5 2.7 N (see Table 1). Maximum polymerization shrinkage forces among the four curing modes were statistically different (p = .001). The maximum force of the EXP mode was statistically lower than MED, 2-STEP, and STD curing modes when Duncans multiple range test was used as a post hoc test. If Tukeys post hoc test was used, the maximum force of the exponential (EXP) mode was statistically lower than only the 2-STEP and STD mode, but statistically similar to MED mode (p = .051). The MED, 2-STEP, and STD modes were statistically equivalent (contained in a second homogeneous subset). Following cessation of light application, the 2-STEP force development

TABLE 1. MAXIMUM RATE AND MAXIMUM FORCE.

Experimental Group*

Mode/Curing Unit Exposure

Maximum Force Rate N/s (SD)

Maximum Force N (SD)

2 4 3 1

EXP Trilight, 40 MED Trilight, 80 STEP Highlight, 40 STD Trilight, 40

0.35 (0.06) 0.33 (0.03) 0.46 (0.07) 0.44 (0.03)

20.4 (2.5) 25.8 (1.0) 27.4 (5.8) 30.5 (2.7)

Vertical lines join statistically equivalent groups *Five specimens per experimental group

332

JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

curve reached an inflection point sooner than the STD specimens (see Figure 3), resulting in a lower maximum force (27.4 N vs. 30.5 N). Whereas the MED force:time curve had a lower initial slope similar to the EXP mode (0.33 N/s vs. 0.35 N/s), the final force of the MED mode was higher (25.8 N vs. 20.4 N) and approximately equal to that of the STEP mode (27.4 N). The EXP mode curve had both a low maximum force rate (N/s) and the lowest maximum polymerization shrinkage force (20.4 N). The EXP mode reduced maximum polymerization shrinkage force to 67% of the maximum value seen using the 40-second STD exposure (20.4 N/ 30.5 N). Real-Time Infrared Spectroscopy Cure mode (or EAS) had no significant influence on DC (p = .0885). However, cure mode did significantly interact with the depth (p = .003). Depth had a significant influence on DC (p = .0001); however, depth also interacted significantly with cure type (p = .003). Therefore, a series of one-way ANOVAs were run to study these effects individually (Figure 4). Both cure mode and depth had a significant influence on conversion rate (p = .0001); however, they interacted significantly (p = .0015). Again, a series of one-way ANOVAs were performed to examine these differences (Figure 5).

Figure 3. Representative force:time curves for Elipar Trilight and Elipar Highlight.

On the top surface, the use of the three different cure methods (standard 40 s = STD, exponential 40 s = EXP, medium 80 s = MED) did not significantly affect the total extent of cure 100 seconds after light initiation (see Figure 4 and Figure 6, B). However, use of the EXP cure mode at this depth significantly reduced conversion rate compared to the two other modes (see Figures 5 and 6, A). The fastest conversion rate was seen using the 40-second STD exposure (9.5%/s), followed by the 80-second MED intensity (7.2%/s), and the lowest was EXP 40-second exposure (5.7%/s). The EXP mode slowed the conversion rate to only 61% of the peak value seen using the 40-second STD exposure (5.7%/s 9.3%/s). However, even with this slowing,

the overall extent of cure was not affected (see Figures 4 and 6, B). At 1.0-mm depth, there was no significant difference in conversion 100 seconds after initiation with respect to the different methods of curing (see Figures 4 and 6, D). However, at 1.0-mm depth cure mode significantly affected peak conversion rate (see Figures 5 and 6, C). Maximal conversion rates, in descending order, were STD (5.8%/s), followed by MED (3.7%/s), which was statistically equivalent to the 40-second EXP exposure value (3.3%/s). Thus, at 1.0-mm depth, 40-second STD exposure had a significantly faster conversion rate than did the EXP or 80-second MED exposure. The 40-second EXP cure mode

VOLUME 12, NUMBER 6, 2000

333

EFFECT OF RAMPED LIGHT INTENSITY ON POLYMERIZATION FORCE AND CONVERSION IN A PHOTOACTIVATED COMPOSITE

produced a conversion rate that was only 57% of that of the 40-second STD (3.3/5.8%/s). At a depth of 2.0 mm from the top surface, for the first time, there was a significant difference in conversion values 100 seconds after light initiation with respect to the different curing methods (see Figures 4 and 6, F). The STD and EXP cure modes produced statistically equivalent conversion values, but the 80second MED intensity conversion was significantly greater only than that of the 40-second EXP method. There was no significant difference in conversion rates among the three methods at 2.0-mm depth from the top surface (see Figures 5 and 6, E).
DISCUSSION

Figure 4. Maximal conversion values measured 100 seconds following light initiation. N = three speciments per group. Vertical bar = 1 standard deviation. Groups with similar letter are statistically equivalent.

Figure 5. Maximal conversion rates with respect to light intensity delivery, exposure duration, and depth from the top surface. N = three specimens per group. Vertical bar = 1 standard deviation. Within a depth, similar lower case letters describe statistically equivalent groups

For low- or variable-intensity lightcuring to serve as a rate-limiting factor on the slope of the force:time curve, the PD must not only exceed the minimal energy to initiate conversion but also fall below the saturation point of the available photoinitiator. Within this intensity range, the maximum rate of cure is proportional to intensity raised to the half power (I0.5).19 Accordingly, doubling the intensity results in only a 1.414-fold increase in conversion rate. As light intensity continues to increase beyond a certain point, a rate saturation effect is reached at which the maximum rate of conversion is hardly changed. This rate limitation is attributable, at least

334

JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

Figure 6. Effect of cure type on rate of conversion (%/s) and degree of conversion (%) at top surface (A and B) 1.0-mm simulated depth from surface (C and D), and 2.0-mm simulated depth from surface (E and F).

VOLUME 12, NUMBER 6, 2000

335

EFFECT OF RAMPED LIGHT INTENSITY ON POLYMERIZATION FORCE AND CONVERSION IN A PHOTOACTIVATED COMPOSITE

partly, to a decrease in the production of initiating radicals under intense illumination.20 Therefore, further increases in light intensity may not affect the rate of conversion. However, within the previously mentioned intensity range, the conversion rate can be manipulated by acting on the initiation rate through the intensity (or PD) of the curing light. The initial low PD of a variable light intensity EAS reduces the formation of free radicals, thereby decreasing the rate of conversion. Thus, reduced intensity values allow an extended time for composite flow prior to mobility restrictions brought on by gelation and solidification of the irradiated material. Higher light intensity following the initial low PD phase of the EAS is necessary to obtain a DC equivalent to a standard single-intensity EAS. This equivalent DC must be achieved without incurring a clinically unacceptable time penalty for the total exposure duration. Slowing the conversion rate by lowering the PD is most critical at the unbound, top surface of the restoration that can relieve stress by flow. The external surface of the restoration also is the most highly irradiated surface. Deeper within the restoration, polymerization rates and shrinkage force rates are lower, since light attenuation by the overlying thickness of composite slows the reaction rate. Therefore, to be effective, a variable-intensity EAS should

demonstrate a reduced rate of conversion (%/s or DC/s) at the top surface. One would also expect a reduced force rate (N/s), an equivalent degree of monomer conversion (DC), and exposure durations roughly equivalent to a standard, single-intensity curing mode. The EXP ramp was successful in reducing the conversion rate at the top surface and at 1.0 mm beneath the top surface, while obtaining a DC that was statistically equivalent to the STD 40-second cure mode. The lower conversion rates (%/s) recorded with EXP and MED cure modes had correspondingly low rates of polymerization shrinkage force development. Whereas a similar DC was observed at the top surface and 1-mm depths, maximum polymerization shrinkage forces recorded with these two EAS were dissimilar. The mean maximum force-rates (N/s) of MED and EXP cure modes were similar; however, the EXP force:time curves had lower inflection points and lower mean maximum force. The lower force rate, lower maximum force, and equivalent DC of the EXP mode would presumably optimize maintenance of the bonded interface without compromising the mechanical properties and biocompatibility that are associated with a high DC. The 80-second MED EAS demonstrated both a reduced conversion rate (%/s) and a reduced maximum

force rate (N/s). The conversion rate using this exposure method was approximately 75% that of the 40-second STD exposure (7.2 %/s 9.5%/s). This reduction in force rate with reduced PD is consistent with previously reported results.9 Lower conversion rate with an accompanying reduction in the force rate at the surface of the composite may be responsible for the improved marginal adaptation reported by several studies that have examined reduced PD curing.7,1012 Halving the intensity, but doubling the exposure time (MED: 400 mW/ cm2 80 s vs. STD: 800 mW/cm2 40 s) to provide similar ED (32 J/cm2) had no effect on total conversion at the top surface. Similarly, there was no difference in degree of conversion at a simulated depth of 1.0 mm 100 seconds after light initiation. Also, at this depth, the 80-second MED intensity conversion rate was equivalent to both the 40-second STD and EXP rates. However, at 2.0 mm from the top surface, the 80-second MED exposure produced a significantly greater DC than was achieved with the EXP mode. This increased DC is probably related to the more extended temperature increase from the curing light and to the exotherm reaction. These factors may result in higher thermal mobility of the reacting network, which, in turn, leads to enhanced conversion of C=C double bonds. Although the DC achieved with the

336

JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

MED mode at 2.0 mm was numerically higher than that with the 40second STD mode, they were statistically equivalent. At 2.0-mm depth, there was no significant difference in conversion rates among the different treatments. The current study confirmed findings that evaluated the soft-start, or stepped light intensity, mode of the Elipar Highlight. A prior study found statistically equivalent force rates, maximum forces, and DC for equivalent exposure durations of the STD and STEP cure modes. The current study used the STD cure mode of the newer Elipar Trilight as a control for STEP cure mode of the earlier Elipar Highlight. As in the prior study, the STD and STEP cure modes resulted in similar force rates and maximum forces. Whereas some studies report improved marginal adaptation using the Elipar Highlight,12,21 the results of this study and a prior study found no difference in force rate or maximum force between the STD and STEP cure modes.22 Such findings suggest that no improvement in marginal adaptation would be expected because of decreased tensile forces with the specific composite and C-factors tested. This is in agreement with Friedl and colleagues,23 who found that using a low-start curing-light intensity did not provide better marginal adaptation in Class V composite resin and polyacid-modified resin restorations.

In the present study, a single lightcuring unit was used to cure a 2.5-mm diameter cylindrical composite specimen from one side only. A prior study tested a higher C-factor (d = 5.0 mm, h = 1.25, C = d/2h = 2.0) where the 5.0-mm diameter sample required simultaneous use of two curing lights, owing to concerns over depth-of-cure and the temporal effects of a stepped power density energy application sequence.22 The total energy flux simultaneously applied from both sides of the sample may have exceeded the ED normally applied in a clinical situation, where only one unbound external surface of the restoration would be irradiated. Therefore, the present study used a single light-curing unit to irradiate a smaller diameter, lower C-factor specimen (d = 2.5 mm, h = 1.25 mm, C = d/2h = 1.0) from only one side. Although the force and force rate in the both of these studies were equivalent for the STD and STEP modes, the values reported by the current study (C = 1.0) are numerically lower than those reported from the prior study (C = 2.0). These findings are in agreement with those of Feilzer and co-workers,15 who found that shrinkage stress is dependent on the ratio of bound to unbound surface area (d/2h) and independent of the volume of a restoration. One would expect a lower maximum force with a numerically lower ratio of bound

to unbound surface area (C-factor) through stress relief by flow from the unbound surface area. It could also be speculated that the difference in energy flux resulting from the simultaneous use of two curing lights in the prior study might have resulted in higher degrees of conversion and, therefore, higher contraction forces. However, the degree of conversion was similar in both studies (DC 60%). Therefore, since the same composite (Pertac II), curing light (Elipar Highlight), and energy application sequence (40-s 2-STEP) was used in both studies, any difference in the maximum shrinkage force between the two studies should be related to the difference in C-factor. The mean maximum force in the present study, with C = 1.0, was 5.6 (1.2) MPa versus the 8.8 (0.6) MPa mean maximum shrinkage force recorded in the prior study, with C = 2.0. Since degree of conversion, rate of conversion, shrinkage force rate, and maximum shrinkage force may be strongly influenced by chemical and physical properties unique to each composite, the magnitude of the reported results would vary with the specific composite brand evaluated. Polymerization shrinkage force generally depends on the composition of the monomers in the matrix, diluents added to improve handling, filler type, filler loading, filler silanization, the resultant modulus of elasticity, and the coefficient

VOLUME 12, NUMBER 6, 2000

337

EFFECT OF RAMPED LIGHT INTENSITY ON POLYMERIZATION FORCE AND CONVERSION IN A PHOTOACTIVATED COMPOSITE

of linear thermal expansion.2429 Additionally, the photoinitiator and accelerator concentrations may affect the rate of conversion (%/s) and final DC (%) that can be expected with a specific light intensity.30 Therefore, the clinician should not assume that a pre-set energy application sequence for photopolymerization is optimal for all composites. Thus, the power density, exposure duration, and energy application sequence of the curing light must be tailored to the overall composition of the composite to maximize the degree of conversion while minimizing the shrinkage stress rate and maximum stress.
CONCLUSION

not be expected to compromise the physical properties of the restorative material, owing to a similar degree of conversion.
DISCLOSURE AND ACKNOWLEDGMENTS

9. Unterbrink G, Muessner R. Influence of light intensity on two restorative systems. J Dent 1995; 23:183189. 10. Bouschlicher MR, Vargas MA, Boyer DB. Effect of composite type, light intensity, configuration factor, and laser polymerization on polymerization contraction forces. Am J Dent 1997; 10:279283. 11. Feilzer AJ, Dooren LH, de Gee AJ, Davidson CL. Influence of light intensity on polymerization shrinkage and integrity of restoration-cavity interface. Eur J Oral Sci 1995; 103:322326. 12. Goracci G, Mori G, deMartinis LC. Curing light intensity and marginal leakage of resin composite restorations. Quintessence Int 1996; 27:355362. 13. Mehl A, Hickel R, Kunzelmann KH. Physical properties and gap formation of light-cured composites with and without softstart-polymerization. J Dent 1997; 25:321330. 14. Burgess JO, De Goes M, Walker R, Ripps AH. An evaluation of four light-curing units comparing soft and hard curing. Pract Periodont Aesthet Dent 1999; 11:125132. 15. Feilzer AJ, de Gee AJ, Davidson CL. Setting stress in composite resin in relation to configuration of the restoration. J Dent Res 1987; 66:16361639. 16. Dlugokinski MD, Caughman WF, Rueggeberg FA. Assessing the effect of extraneous light on photoactivated resin composites. J Am Dent Assoc 1998; 129:11031109. 17. Rueggeberg FA, Caughman WF, Chan DCN. Novel approach to measure composite conversion kinetics during exposure with stepped or continuous light-curing. J Esthet Dent 1999; 11:197205. 18. Rueggeberg FA, Hashinger DT, Fairhurst CW. Calibration of FTIR conversion analysis of contemporary dental resin composites. Dent Mater 1990; 6:241249. 19. Billmeyer F Jr. Textbook of polymer science. 3rd Ed. New York: John Wiley and Sons, 1984:4958. 20. Decker C, Decker D, Morel F. Light intensity and temperature effect in photoinitiated polymerization. In: Scranton AB, Bowman CN, Peiffer RW, eds. Photopoly-

This study was supported by a grant from ESPE America. The authors thank John M. Winterbottom for programming and operation of the MTS 318 Biaxial Test System.
REFERENCES

1. Rueggeberg FA, Caughman WF, Curtis JWJ, Davis HC. A predictive model for the polymerization of photo-activated resin composites. Int J Prosthodont 1994; 7:159166. 2. Eick JD, Welch FH. Polymerization shrinkage of posterior composite resins and its possible influence on postoperative sensitivity. Quintessence Int 1986; 17:103111. 3. Soltesz U, Benkeser G. Fatigue behaviour of filling materials. In: Kawahara H, ed. Oral implantology and biomaterials. The Netherlands: Elsevier Science Publishers, 1989:281286. 4. Sidhu SK, Henderson LJ. Dentin adhesives and microleakage in cervical resin composites. Am J Dent 1992; 5:240243. 5. Bowen RL, Rapson JE, Dickson G. Hardening shrinkage and hygroscopic expansion of composite resins. J Dent Res 1982; 61:654658. 6. Lutz F, Krejci I, Barbakow F. Quality and durability of marginal adaptation in bonded composite restorations. Dent Mater 1991; 10:107113. 7. Kanca J, Suh BI. Pulse activation: reducing resin-based composite contraction stresses at the enamel cavosurface margins. Am J Dent 1999; 12:107112. 8. Uno S, Asmussen E. Marginal adaptation of a restorative resin polymerized at reduced rate. Scand J Dent Res 1991; 99:440444.

Maximum polymerization shrinkage force and maximum rate of force development were significantly lower with ramped light intensity (EXP 40-s Trilight). Intermediate intensity (MED 80-s Trilight) also resulted in a significantly lower rate of force development than standard (STD) or stepped intensity (STEP) light-curing. Ramped light intensity slowed the rate of cure at the top surface and at 1.0 mm beneath the surface, but did not significantly affect the total conversion value compared to a standard 40-second, single-intensity cure mode. The slower conversion rate resulting from ramped light intensity helped to reduce the rate and maximum polymerization stress, but would

338

JOURNAL OF ESTHETIC DENTISTRY

BOUSCHLICHER AND RUEGGEBERG

merization: fundamentals and applications. Vol. 673. Washington, DC: American Chemical Society, 1997:6380. 21. Mehl A, Manhart J, Kremers L, Kunzelmann KH, Hickel R. Physical properties and marginal quality of Class II composite fillings after softstart polymerization. J Dent Res 1996; 76:279. (Abstr) 22. Bouschlicher MR, Rueggeberg FA, Boyer DB. Effect of stepped light intensity on polymerization force and conversion in a photoactivated composite. J Esthet Dent 2000; 12:2332. 23. Friedl KH, Schmalz G, Hiller K, Mrkl A. Marginal adaption of Class V restorations with and without softstart-polymerization. Oper Dent 1999; 25:2632.

24. Asmussen E. Composite restorative resins: composition versus wall-to-wall polymerization contraction. Acta Odontol Scand 1975; 33:337343. 25. Condon JR, Ferracane JL. Assessing the effect of composite formulation on polymerization stress. J Am Dent Assoc 2000; 131:497503. 26. Condon JR, Ferracane JL. Reduction of composite contraction stress through nonbonded microfiller particles. Dent Mater 1998; 14:256260. 27. Aarnts MP, Akinmade A, Feilzer AJ. Effect of filler load on contraction stress and volumetric shrinkage. J Dent Res 1999; 78:482. (Abstr)

28. Sakaguchi RL, Ferracane JL. Stress transfer from polymerization shrinkage of a chemical-cured composite bonded to a pre-cast composite substrate. Dent Mater 1998; 14:106111. 29. Versluis A, Douglas WH, Sakaguchi RL. Thermal expansion coefficient of dental composites measured with strain gauges. Dent Mater 1996; 12:290294. 30. Aarnts MP, Akinmade A, Feilzer AJ. Effects of photoinitiator concentration on conversion, strength, and stress development. J Dent Res 1999; 78:234. (Abstr)

Reprint requests: Murray R. Bouschlicher, DDS, MS, Department of Operative Dentistry, College of Dentistry, The University of Iowa, Iowa City, IA 52242; e-mail: murray_bouschlicher@uiowa.edu 2000 BC Decker Inc

VOLUME 12, NUMBER 6, 2000

339

Polymerization Depth of Contemporary Light-Curing Units Using Microhardness


FREDERICK A. RUEGGEBERG, DDS, MS* JANET W. ERGLE DONALD J. METTENBURG

ABSTRACT

Purpose: This research investigated composite depths of cure using a variety of light-curing units and exposure protocols. Materials and Methods: Composite (Herculite XRV, shade A2, Kerr, Orange, California) was exposed in opaque compules to conventional quartz tungsten halogen (QTH) units, soft-start units, high-intensity QTH and plasma arc curing (PAC) lights, and one argon laser. Cured compules were sonicated to remove uncured composite and were sectioned and polished along the long axis to expose cured composite. Knoop hardness was measured 0.5 mm from the irradiated, top surface and then at 1.0 mm and in 1.0-mm increments until reliable readings could no longer be obtained. Hardness values were compared by analysis of variance at similar depths within a specific curing-light classification, using the hardness of the standard 40-second conventional QTH exposure as comparison (Dunnetts t-test). Depth of cure was defined as the deepest hardness value found equivalent to that at 0.5-mm depth for a specific curing light and scenario. Results: Conventional QTH lights provided similar hardness profiles. At 2-mm depth, use of a different unit or curing tip made no difference in hardness compared with the standard. At this depth, soft-start (pulse-delay and step-cure) methods yielded similar hardness as the standard. High-intensity QTH lights provided similar hardness at 2-mm depth in 10 seconds as the standard 40-second exposure. Plasma-arc exposure for less than 10 seconds produced inferior hardness compared with the standard. A 10-second PAC and a 5-second laser exposure gave equivalent hardness at 2-mm depth as the 40-second standard. Depth of cure for almost all curing scenarios was not greater than 2 mm.
CLINICAL SIGNIFICANCE

Similar-type conventional QTH lights with different tip diameter (8 and 12 mm) provide similar composite cure characteristics. Soft-start techniques provide similar cure profiles as conventional QTH technique when used according to manufacturers recommendations. High-intensity QTH units and the argon laser can reduce exposure time while providing composite with similar hardness as conventional QTH curing. Plasma-arc exposure should be at least of 10 seconds duration to provide hardness equivalent to that achieved with conventional 40-second QTH exposure. Even with consideration of high-intensity curing units, composite increments should still be no greater than 2 mm to provide homogeneous hardness. (J Esthet Dent 12:340349, 2000)

*Professor and Section Director, Research Assistant, Dental Materials Section, Department of Oral Rehabilitation, School of Dentistry, Medical College of Georgia, Augusta, Georgia 340
JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG ET AL

or years, conventional methods of exposing photoinitiated resin composite materials have been with relatively low-powered (3575 W) quartz tungsten halogen (QTH) units.1 These units provide a wide spectral output, after filtering, between 400 and 500 nm25; however, the output intensity is typically lower than 1000 mW/cm2. Conventional QTH curing sources operate in a continuous output mode; when activated, the source is driven to provide full output intensity for the selected exposure duration.6 When using these light sources, recommended exposure duration for 2-mm incremental composite placement is between 40 and 60 seconds.79 Currently, a wide variety of curing lights and methods is available.6 These newer units offer remarkably higher output intensity levels than conventional QTH units previously described. Also, many features related to output control during exposure have been provided. High-intensity QTH lights are a separate classification of curing units. Typically, the intensity of these sources is in excess of 1000 mW/cm2.6 This level could be reached merely by replacing the conventional light-curing tip with one that concentrates light into a smaller distal tip end diameter. These tips, referred to as turbo tips, operate by gathering light from a large proximal tip diameter near the focal plane of the QTH source

and concentrating that light into a smaller distal tip diameter.10 This concentration results from use of a tapered fiber optic glass bundle. The source voltage driving a conventional QTH bulb can also be increased to provide increased radiant output.11 This mechanism, along with a turbo tip, is used in a popular unit to provide enhanced output level: Optilux 501 in burst mode (Demetron Research, Danbury, Connecticut). An additional method of providing enhanced output levels from a QTH light is use of a high wattage-source generating light remote from the operating field. The intense irradiation from this unit is directed to the restoration via means of a liquid-filled light guide (Kreativ Kuring Unit, Kreativ Inc., San Diego, California). The rate of composite cure, and the development of stress from polymerization shrinkage at the tooth resin interface as well as within the tooth structure itself, are clinical concern.1217 To minimize shrinkage, manufacturers have introduced QTH-based lights specifically designed to provide low levels of intensity during the initial exposure.18,19 Lowering light intensity, is thought to reduce the rate at which free radicals are formed and the rate at which polymerization occurs. Providing a low initial rate of cure is intended to allow composite to flow from the unbonded surfaces to relieve stresses prior to

attaining a gel point.15,20 These systems use a number of methods to provide control over light output.6,21 Step-cure units generate an initial low-level output for a predetermined time into the curing exposure. Immediately following this phase, the unit generates full output for the remainder of the exposure. The initial light intensity level can also be slowly increased during the exposure and then maintained at a high, fixed level. Such a modification of light output is termed ramping. A modification of the step-cure method is the pulse-delay procedure.20 In this method, a shortduration, low-intensity exposure is provided only to the last incremental layer. A designated time is allowed to pass, during which the composite surface is contoured and polished. At completion of this time, a second, more intense and longer exposure is provided to finalize conversion. A completely different method of providing extremely high levels of irradiance uses the newly introduced plasma arc curing (PAC) units.6 Output levels from these sources are remarkably higher than those of conventional or high-intensity QTH units. The PAC lights also use liquid light guides to reduce levels of infrared and ultraviolet radiation. Manufacturers claim these sources can effectively reduce clinical exposure duration to only 1 to 10 seconds, whereas a 40-second

VOLUME 12, NUMBER 6, 2000

341

POLYMERIZATION DEPTH OF CONTEMPORARY LIGHT-CURING UNITS USING MICROHARDNESS

exposure from a conventional QTH light would be required. The argon ion laser was once popular for curing photoactivated composites.22 However, fewer manufacturers are supporting this technology for dental curing purposes. Many claims have been made related to this curing mode, indicating significant reduction in clinical exposure time compared to the conventional QTH methods while yielding similar conversion and physical properties.22,23 Thus, claims of reduced exposure requirements while providing similar conversion values and depths of cure as conventional QTH units are made. Clinical success of a photocured dental restoration may depend upon obtaining high levels of conversion.24 With the clinicians reputation and livelihood being based on the success of restorations placed, manufacturers claims of reduced exposure time and maintenance of high conversion levels of these new light-curing units and methods need to be verified. The purpose of this research is to examine the depths of cure, using microhardness values, when a variety of commercial curing units and methods are used to polymerize a single lot of commercial, photoactivated dental composite. The hypothesis tested is that the hardness value obtained using a conventional QTH exposure method

(standard) is equivalent to those achieved when the other curing units and methods are applied as directed by the manufacturer.
MATERIALS AND METHODS

The composite used throughout this study was a hybrid, photoactivated material (Herculite XRV, shade A2, lot 712557, Kerr Dental Products, Orange, California). To minimize light penetration from reflection and scattering, composite was cured in black, opaque, plastic compules. This method produced a stringent environment in which to examine specifically the effect of irradiated light without the confounding factors of side reflection or translucency of surrounding matrix structure. Preparation for specimen curing included removal of the curved compule end with a sharp knife blade and removal of the plunger from the other end. A small piece of Mylar was placed on a tabletop, and the open compule end, where the plunger had been, was placed on top of the plastic square. The composite was then condensed against the Mylar surface, transferring the uncured mass from one compule end to the other. The packed compule was positioned in a holding jig that aligned the specimen vertically, with the top, Mylar-covered end facing upward (Figure 1). This jig was attached to the platform of a laboratory jackstand, allowing for vertical movement of the specimen. The distal end of a light-curing tip was held in

position so that the tip center was coincident with the compule long axis. Composite was exposed with the curing tip end held 1 mm from the Mylar surface. Fifteen minutes following exposure, the compule was placed in solvent (methyl ethyl ketone, M209-4, Fisher Scientific, Fair Lawn, New Jersey) and sonicated for 1 hour. Immersion and sonication removed uncured composite by dissolution. The specimen was then placed in another holding device that held the compule along its horizontal axis. The compule was ground in this device, to remove the plastic casing and to expose cured composite along a topbottom plane. The specimen was ground almost in half, preserving mechanical retention in the holder and compule. Attention was given to slow speed and to wet grinding to ensure a minimal impact of specimen preparation on hardness readings. The sequence of abrasives used on the rotary grinding apparatus (Ecomet III, Buehler Corp., Lake Bluff, Illinois) was as follows: 240, 320, 400, and 600 grit, wet-ground (silicon carbide, Leco Corp., St. Joseph, Michigan); diamond paste (Nos. 15, 6, and 1 synthetic diamond paste, Lay Industrial Diamond Corp., Deerfield Beach, Florida) on a polishing cloth (No. 40-7628, Buehler Corp.). Following polishing, specimens were prepared for microhardness testing. Each compule was positioned in another holding jig that

342

JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG ET AL

at a similar depth, and then applying Dunnetts two-tailed post hoc t-test, using the standard value as control. The depth of cure of each light-curing treatment was also determined. This value was defined as the greatest depth that yielded hardness value statistically equivalent to that seen at the 0.5-mm depth for a specific treatment. For this analysis, a one-way ANOVA was performed using Dunnetts one-tailed t-test with the 0.5-mm hardness value used as control. All statistical testing was performed at a pre-set alpha of 0.05.
Figure 1. Schematic representation of specimen light-curing.
RESULTS

maintained the polished surface parallel with the mounting stage of a Knoop hardness testing machine (model MO, Tukon Tester, ACCO Wilson Instrument Division, New York, New York). Hardness readings were obtained using a 0.5-kg load and 10 times magnification. Triplicate readings were made at a distance of 0.5 mm from the top, irradiated surface. Readings were averaged to represent the hardness of that specimen at the specified depth from the top surface. Hardness values were also obtained at a depth of 1.0 mm and then at 1.0-mm increments until values could no longer be validly determined. The light-curing units and modes of photopolymerization used represented a wide variety of contemporary models and methods (Table 1).

Where possible, all units were used with new light sources. For comparative purpose, the hardness obtained when using the conventional QTH unit (Optilux 401) with an 8-mm diameter tip and a 40-second exposure were considered to represent the standard. This unit type and exposure duration were believed to represent the majority of conventional light-curing units in use. Three specimens were made for each test condition (72 specimens). Statistical analysis consisted of comparing the hardness at each specified depth from the top surface using the different light sources and curing modes to that obtained using the standard method previously described. This analysis was accomplished by performing a one-way analysis of variance (ANOVA) on hardness values

Conventional Quartz Tungsten Halogen Units Figure 2, A, presents hardness values at different depths for conventional QTH units. Note the data line representing the standard curing mode (Optilux 401, 8-mm diameter tip, 40-s exposure). This curve is seen in all subsequent hardness profile graphs and represents hardness values obtained using the standard QTH unit and exposure duration. The hardness values of all QTH units within each depth were compared to the values obtained for the standard. At depths of 0.5, 1.0, and 2.0 mm, the standard treatment yielded hardness values equivalent to those for all other units. At 3-mm depth, hardness using the standard treatment was equivalent to values from all other units with the exception of those from the Optilux 401 with turbo

VOLUME 12, NUMBER 6, 2000

343

POLYMERIZATION DEPTH OF CONTEMPORARY LIGHT-CURING UNITS USING MICROHARDNESS

TABLE 1. LIGHT-CURING UNITS USED IN THE PRESENT STUDY.

Classification of Curing Unit

Tip Diameter (mm)

Exposure Duration (s)

Manufacturer

Serial Number

Conventional QTH Optilux 401 Optilux 401 Optilux 500 Optilux 501 Optilux 501 Soft-start QTH Highlight Highlight VIP High-intensity QTH Optilux 501 Kreativ Kuring Unit Kreativ Kuring Unit Kreativ Kuring Unit Kreativ Kuring Unit Plasma arc Apollo 95E Apollo 95E Apollo 95E Apollo 95E ADT 1000 PowerPAC Virtuoso ARC Light Argon ion laser Accucure 3000 Accucure 3000 Accucure 3000

8 Turbo tip (14 10) 8 8 12 10 10 12 Turbo tip (14 8) 8 8 8 8 470 nm tip, 7 470 nm tip, 7 470 nm tip, 7 470 nm tip, 7 8 8 6 8 12 12 12

40 20 40 40 40 40 40 step* Pulse-delay 10, burst 10, normal 10, ramp 10, boost 20 1 2 3 Step-cure 10 10 11 10 5 10 20

Demetron/Kerr, Danbury, CT Demetron/Kerr Demetron/Kerr Demetron/Kerr Demetron/Kerr ESPE America, Norristown, PA ESPE America Bisco, Schaumburg, IL Demetron/Kerr Kreativ Inc., San Diego, CA Kreativ Inc. Kreativ Inc. Kreativ Inc. DMD, Woodland Hills, CA DMD DMD DMD ADT, Corpus Christi, TX ADT Den-Mat, Santa Maria, CA Air Techniques, Hicksville, NY LaserMed, Salt Lake City, UT LaserMed LaserMed

4733101 4733101 5800112 5810289 5810289 3801548 3801548 013-205 5810289 6801501520 6801501520 6801501520 6801501520 NE 808 349 NE 808 349 NE 808 349 NE 808 349 1126 5333 99F1634A760 5406 9803896 9803896 9803896

*10 s at low level, 30 s at continuous output; 200 mW/cm2 for 3 s, 3-min wait, exposure to 600 mW/cm2 for 10 s; 2 s at 150 W, 4 s at 300 W. QTH = quartz tungsten halogen (the standard); DMD = Dental Medical Diagnostics; ADT = American Dental Technologies.

tip for 20-second exposure, which were lower. At 4- and 5-mm depths, hardness values using the standard method were equivalent to those from all other units. The depths of cure for the various treatments (as previously defined) were as follows: standard unit (Optilux 401, 8-mm tip, 40 s) and Optilux 501, 8-mm tip, 40-s exposure, 2 mm; Optilux 401 with turbo tip, 20-second

exposure, Optilux 500, 8-mm tip, 40-second exposure, and Optilux 501 12-mm tip, 40-second exposure, 1 mm. Soft-Start Quartz Tungsten Halogen Units Hardness values from the top surface for soft-start QTH units were equivalent (within a specific depth) among the standard method and all

soft-start units up to a depth of 2 mm (Figure 2, B). Thereafter, hardness of the pulse-delay technique was significantly less than than that of other methods. At 5-mm depth, the hardness value for the standard method was lower than that for either the continuous or step output from the Highlight (ESPE America, Norristown, Pennsylvania). Depth of cure for the standard method

344

JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG ET AL

was 2 mm, whereas for all methods from the Highlight, cure depth was 3 mm, and for the pulse-delay technique, it was only 1 mm. High-Intensity Quartz Tungsten Halogen Units Hardness values at a depth of 0.5 mm from the top irradiated surface for high-intensity QTH units

were equivalent to standard, except the 10-second normal and ramped exposures of the Kreativ unit, which were less (Figure 2, C). At 1.0- and 2.0-mm depths, all hardness values were equivalent to those of the standard. At 3-mm depth, only the Kreativ 10-second boost and 20-second normal exposures provided hardness values equivalent

to those of the standard; all others were lower. Hardness of Kreativ 20-second normal exposure and standard unit was equivalent at 4-mm depth, whereas only the standard unit provided readable values at 5-mm depth. The standard treatment provided a depth of cure of 2 mm, whereas the Optilux 501 with turbo tip used in the 10-second

D Figure 2. Hardness profile ( 1 SD) of composite exposed: A, using various quartz tungsten halogen units and tip diameters (all Optilux models); B, using various soft-start, quartz tungsten halogen units; C, using various high-intensity quartz tungsten halogen units; D, using various plasma arc (PAC) units; and E, using various exposure durations from an argon ion laser (n = 3 specimens/group).

VOLUME 12, NUMBER 6, 2000

345

POLYMERIZATION DEPTH OF CONTEMPORARY LIGHT-CURING UNITS USING MICROHARDNESS

burst mode had a depth of cure of less than 1 mm. The Kreativ unit, used in all four scenarios, provided similar 2.0-mm depths of cure. Plasma Arc Curing Units Hardness values at 0.5-mm depth for PAC units and exposure methods were equivalent to standard, with the exception of all specimens exposed using the Apollo 95E (Dental Medical Diagnostics, Woodland Hills, California), which yielded lower values (Figure 2, D). This trend was also seen at 1.0 mm deep, except that at this level, the hardness from a 3-second Apollo 95E exposure was equivalent to that of standard. At 2-mm depth, no hardness values for the 1-second Apollo 95E could be recorded. Only the PAC units providing at least 10-second exposure (Virtuoso, ARC Light, ADT 1000 [American Dental Technologies, Corpus Christi, Texas], and the PowerPAC units [American Dental Technologies]) demonstrated hardness values similar to those of the standard. The only exposure scenario using the Apollo 95E that provided recordable hardness values at 3-mm depth was the step-cure mode. This technique, as well as use of the ARC Light (Air Techniques, Hicksville, New York), yielded hardness values lower than all other units as well as the standard method at 3-mm depth. At 4-mm depth, the ADT 1000 and Virtuoso units had hardness values equivalent to those of

the standard method, but those from the PowerPAC were lower. The only hardness readings obtainable at 5-mm depth were from the standard technique and from the Virtuoso unit, which were equivalent. The standard mode provided a depth of cure of 2 mm. Of the PAC lights providing exposure duration of at least 10 seconds, the ADT 100 unit yielded a depth of cure of 3 mm, the Virtuoso 2 mm, and the PowerPAC and ARC Lights gave 1-mm cure depths. A 1- or 2-second exposure and the step-cure mode from the Apollo 95E unit gave a depth of cure of 1 mm, whereas the 3-second exposure cure depth was less than 1 mm. Argon Ion Laser All hardness values at a depth of 5 mm for an argon ion laser at highest output (250 mW) using various exposure durations were equivalent, but lower than values achieved by the standard method (Figure 2, E). However, at depths of 1.0 and 2.0 mm, all hardness values from the laser were equal to those of the standard method. At 3.0-mm depth, hardness from the 5-second laser exposure could no longer be recorded. At this depth, the 10-second laser exposure gave lower hardness values than either the 20-second laser or the standard mode. The only hardness values attainable at 4-mm depth were from the standard technique and the 20-second laser, which were

equivalent. The standard technique provided a 2-mm depth of cure, as did the 5- and 10-second laser exposures. The depth of cure using a 20-second laser exposure was 3 mm.
DISCUSSION

For the conventional QTH units, all lights and methods provided equivalent hardness values within a given depth up to 2 mm, validating the research hypothesis. The reason that the Optilux 401 with 8-mm tip should provide hardness values superior to those achieved by either the 500 or the 501 with similar sized tip and equivalent exposure duration is not known. Perhaps, in the specific Optilux 401 unit tested, the focal plane of the QTH source was aligned better with respect to the proximal curing tip aperture than in the other units. All units had similar type sources (80W Optibulb), so hardness differences may be related to the specific optics of the source and curing tip combinations. Statistically, not all the units provided depths of cure up to 2 mm with respect to hardness values at 0.5-mm depth. The standard unit and method and the Optilux 500 and 501 units with 8-mm tip for 40-second exposure yielded 2-mm cure depths, whereas use of the turbo tip in the Optilux 401 provided only a 1-mm depth. The poor performance of the turbo tip could be related to a focusing effect. The distal end of all curing tips was held at a distance of 1.0 mm from the Mylar

346

JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG ET AL

film during exposure. If the turbo tip had a focal point, the tip could have been used outside of its most effective target-source distance. All soft-start methods provided hardness up to a depth of 2 mm, equivalent to that achieved with the standard technique, validating the research hypothesis. However, past this depth, the pulse-delay technique demonstrated inferior curing. However, it should be noted that this technique is recommended only for the last 1.5- to 2-mm thick increment. Use of a step-cure method of exposure duration equivalent to a continuous one (Highlight step and normal) produced hardness values equivalent to not only each other but also to that of the standard method. These results indicate that the step-cure method would provide a cure value equivalent to that of a continuous exposure, eliminating the possibility of under-cure owing to the use of a lower initial intensity. For the high-intensity QTH units, not all exposure scenarios from the Kreativ or from the Optilux 501 in burst mode with turbo tip provided values equivalent to those achieved with the standard unit, invalidating the research hypothesis. It is interesting to note that no high-intensity QTH unit provided an enhanced depth of cure greater than that seen when using the standard mode, which was 2 mm.

The research hypothesis for PAC lights was not proven true for all exposure scenarios. Those PAC units providing exposure duration of at least 10 seconds generated hardness values similar to those of the standard technique. At 2-mm depth, all PAC lights providing at least this duration had hardness equivalent to that of the standard method. This result demonstrates that a 10-second PAC light exposure provides equivalent composite hardness when a 2-mm increment is used. At this depth, none of the exposure scenarios from the Apollo 95E provided hardness similar to the former. Thus, short exposure duration (3 seconds or less, or a step-cure exposure of less than 6 seconds) from a PAC light is seen not to provide a polymer matrix of properties similar to that of a conventional QTH unit. Only one PAC unit (ADT 1000) indicated a depth of cure value exceeding the 2-mm level of the standard method. Thus, as a rule, PAC lights do not tend to provide enhanced depths of cure when used for 10-second exposures. When used for less than this duration, the depths of cure are low. The research hypothesis for hardness values obtained with the laser was validated. As exposure duration from this unit increased, the depth to which hardness values could be obtained also increased. The 20-second argon exposure mirrored the hardness profile of the 40-second

standard technique. Thus, it is believed that the argon laser can be used to reduce chairside curing time, but perhaps by only one-half of the recommended exposure time recommended when using conventional QTH units. Depths of cure using the laser were equivalent to that of the standard method, with the 20-second exposure providing a cure depth of 3 mm. There was great variation in hardness profiles among the various units and methods compared to the standard QTH mode. In general, use of the different conventional QTH units yielded similar hardness profiles. Use of the turbo tip with this type of unit did reduce exposure duration by half while providing equivalent hardness values at 2-mm depth. However, it did not enhance hardness values above the value achieved with the standard method. The purpose of using highoutput light sources is to reduce chairside exposure time while producing properties equivalent to those that would have been realized if the longer exposure, conventional QTH mode had been used. In many cases, such a reduction of time, while providing equivalent hardness, was realized. All methods of providing high-intensity QTH light provided similar hardness values at 2-mm depth compared to the standard 40-second treatment. Thus, use of the Optilux 501 in Burst mode with a turbo tip for 10-second dura-

VOLUME 12, NUMBER 6, 2000

347

POLYMERIZATION DEPTH OF CONTEMPORARY LIGHT-CURING UNITS USING MICROHARDNESS

tion as well as the Kreativ unit, even in its normal 10-second exposure mode, provided hardness similar to that achieved with the standard method at 2-mm depth. For the PAC lights, the results indicate that a minimum 10-second exposure should be used to provide composite hardness at 2-mm depth similar to that achieved with the standard QTH mode. Exposures of 1, 2, or 3 seconds, or even a step-cure with a PAC light lasting approximately 6 seconds, provided inferior hardness at 2-mm depth compared to all other PAC treatments. A 5-second exposure from the argon ion laser gave hardness values at 2-mm depth equivalent to those with the standard QTH mode. However, the 20-second laser exposure almost perfectly matched the hardness profiles of the standard QTH unit up to a depth of 4 mm. The soft-start methods evaluated were a step-cure and a pulse-delay technique. When considering hardness values at 2-mm depth, both of these methods provided hardness values equivalent to those seen with the standard mode. Thus, clinicians need not fear under-curing composite when using these curing scenarios when the units are used according to the manufacturers directions. It is interesting to note that hardness values for curing scenarios providing light intensity tend to drop dramatically at thickness greater than 2 mm. It was anticipated that use of high-powered curing lights (QTH,

PAC, and the argon laser) would provide cure depths in excess of the currently advocated 2-mm level. However, for most systems, the recommendation of incremental thickness layering to a maximum of 2 mm still holds.79 The results of this research need to be considered with respect to certain factors. First, the depth of cure values presented are with respect to hardness values obtained at the 0.5-mm depth. Such readings may be misleading. If a curing treatment produced a relatively poor hardness at 0.5-mm depth, and also yielded low values throughout the hardness profile, the reader may mistakenly consider that this particular unit yielded a high depth of cure. However, the absolute hardness values could be below those of more efficient curing units. Also, only one brand of composite was tested in this research. Hardness values may not be used as absolute indicators of composite cure, only relative ones. Within a single composite system, comparison of hardness may be used to indicate the relative cure of the resin matrix. However, hardness values cannot be compared between or among other commercial products, because the resin systems and filler content are not similar. It is for this reason that a single composite system was used throughout this study. Other types of composites that may be more or less photosensitive may provide different curing patterns with the

lights tested. Thus, the reader is cautioned in applying the results of this study universally to all types of composites and curing units. However, within the restrictions imposed by the experimental methods, comparisons of the potential for resin cure among different types of curing lights and methods are presented.
CONCLUSIONS

Within the restrictions imposed in the current study, the following conclusions may be made: 1. Use of many different types of conventional QTH curing units and tips did not significantly affect hardness readings compared to the standard curing method at a depth of 2 mm from the irradiated surface. 2. Soft-start QTH curing modes (step-cure and the pulse-delay) provided hardness values at 2-mm depth similar to those of the standard QTH curing method. 3. Use of high-intensity QTH units (stated output level in excess of 1000 mW/cm2) provided hardness values at 2-mm depth equivalent to that of the standard QTH method. Exposure durations of these high-intensity lights may be as low as 10 seconds to provide hardness similar to that with a 40-second exposure using a conventional QTH source. 4. Exposure from PAC lights should be at least 10 continuous seconds to provide hardness

348

JOURNAL OF ESTHETIC DENTISTRY

RUEGGEBERG ET AL

values at 2-mm depth similar to those seen with the standard QTH method. Duration values from 1 to 3 seconds, or use of a step-cure mode, does not provide hardness values equivalent to that achieved with standard QTH methods. 5. The argon-ion laser with a 5-second exposure at 2-mm depth provides hardness values similar to those achieved with the standard QTH 40-second exposure. Use of a 20-second laser exposure yielded hardness profiles similar to those with the standard technique to a depth of 4 mm. 6. Incremental layers of 2-mm thickness are still recommended when using soft-start or highintensity curing methods (QTH, PAC, argon laser), to provide uniform hardness throughout the cured mass.
ACKNOWLEDGMENTS

3. Blankenau RJ, Kelsey WP, Cavel WT, Blankenau P. Wavelength and intensity of seven systems for visible light-curing composite resins: a comparison study. J Am Dent Assoc 1983; 106:471474. 4. Stokes AN, Hodgkinson IJ. Spectral distribution and relative irradiance of dental white-light sources. N Z Dent J 1984; 80:1920. 5. Lee SY, Chiu CH, Boghosian A, Greener EH. Radiometric and spectroradiometric comparison of power outputs of five visible light-curing units. J Dent 1993; 21:373377. 6. Rueggeberg FA. Contemporary issues in photo-curing. Compend Cont Educ Dent 1999; 20(Suppl):S4S15. 7. Council on Dental Materials, Instruments, and Equipment. Visible light-cured composites and activating units. J Am Dent Assoc 1985; 110:100102. 8. Rueggeberg FA, Caughman WF, Curtis JW Jr. Effect of light intensity and exposure duration on cure of resin composite. Oper Dent 1994; 19:2632. 9. Maris JT, Dannheimer MF, Germishuys PJ, Borman JW. Depth of cure of lightcured composite resin with light-curing units of different intensity. J Dent Assoc S Afr 1997; 52:403407. 10. Curtis JW, Rueggeberg FA, Lee AJ. Curing efficiency of the turbo tip. Gen Dent 1995; 43:428433. 11. Fan PL, Wozniak WT, Reyes WD, Stanford JW. Irradiance of visible light-curing units and voltage variation effects. J Am Dent Assoc 1987; 115:442445. 12. Kinomoto Y, Torii M, Takeshige F, et al. Comparison of polymerization contraction stresses between self- and light-curing composites. J Dent 1999; 27:383389. 13. Losche GM. Marginal adaptation of Class II composite fillings: guided polymerization vs reduced light intensity. J Adhes Dent 1999; 186:388391. 14. Burgess JO, DeGoes M, Walker R, Ripps AH. An evaluation of four light-curing units comparing soft and hard curing. Pract Periodont Aesthet Dent 1999; 11:125132.

15. Koran P, Kurschner R. Effect of sequential versus continuous irradiation of a lightcured resin composite on shrinkage, viscosity, adhesion, and degree of polymerization. Am J Dent 1998; 11:1722. 16. Bouschlicher MR, Vargas MA, Boyer DB. Effect of composite type, light intensity, configuration factor and laser polymerization on polymerization contraction forces. Am J Dent 1997; 10:8896. 17. Feilzer AJ, Dooren LH, de Gee AJ, Davidson CL. Influence of light intensity on polymerization shrinkage and integrity of restoration-cavity interface. Eur J Oral Sci 1995; 103:322326. 18. Rueggeberg FA, Caughman WF, Chan DCN. Novel approach to measure composite conversion kinetics during exposure with stepped or continuous light-curing. J Esthet Dent 1999; 11:197205. 19. Light-curing units. Farah JW, Powers JM, editors. The Dental Advisor 1999; 16:15. 20. Kanka J 3rd, Suh BI. Pulse activation: reducing resin-based composite contraction stresses at the enamel cavosurface margins. Am J Dent 1999; 12:107112. 21. Alberts HF. Resin polymerization. Adept Report 2000; 6:3. 22. Blankenau RJ, Kelsey WP, Powell GL, Shearer GO, Barkmeier WW, Cavel WT. Degree of composite resin polymerization with visible light and argon laser. Am J Dent 1991; 4:4042. 23. Kelsey WP, Blankanau RJ, Powell GL, Barkmeier WW, Stormberg EF. Power and time requirements for use of the argon laser to polymerize composite resins. J Clin Laser Med Surg 1992; 10:273278. 24. Ferracane JL, Mitchem JC, Condon JR, Todd R. Wear and marginal breakdown of composites with various degrees of cure. J Dent Res 1997; 76:15081516.

The authors thank the manufacturers of the light-curing units used in this study. Kerr/Sybron is recognized for donation of the composite.
REFERENCES

1. McCabe JF, Carrick TE. Output from visible light-activated units and depth of cure of light-activated composites. J Dent Res 1989; 68:15341539. 2. Cook WD. Spectral distributions of dental photopolymerization sources. J Dent Res 1982; 61:14361438.

Reprint requests: Frederick Rueggeberg, DDS, MS, Department of Oral Rehabilitation, School of Dentistry, Room AD-3265A, Medical College of Georgia, Augusta, GA 30912-1260; e-mail: frueggeb@mail.mcg.edu 2000 BC Decker Inc

VOLUME 12, NUMBER 6, 2000

349

S-ar putea să vă placă și