current therapeutic options for endodontic biofilms

20
Current therapeutic options for endodontic biofilms MARKUS HAAPASALO & YA SHEN Microbial biofilms in the infected root canal space are the primary cause of apical periodontitis. Root canal treatment therefore aims to either remove the biofilms from the root canal or kill all of the microbes in the biofilms. Instrumentation mechanically removes or disrupts biofilm organization and creates sufficient space in the canal to allow effective irrigation and disinfection to occur. While none of the mechanical or chemical factors alone can predictably eradicate the infective agents, their combined action under optimal efforts is the key factor for long-term success of endodontic treatment and healing of the lesion. In this article, the role and impact of various mechanical (physical) and chemical means of attacking root canal biofilms are discussed in light of relevant literature (Fig. 1). Received 14 November 2011; accepted 30 June 2012. Building blocks for success in endodontic treatment Apical periodontitis (AP) is an inflammatory reaction of periradicular tissues caused by a microbial infection in the root canal (1,2). Because the bacteria in the necrotic root canal system grow mostly in sessile bio- films, the success of endodontic treatment depends on effective elimination of such biofilms. The necessary elements in the control of endodontic infection are host defense, instrumentation and irrigation, locally used intracanal medicaments between appointments, root canal filling, and coronal restoration (3,4). Chemomechanical preparation has been regarded as the key element of endodontic treatment (5,6). Impor- tantly, mechanical canal preparation supports disinfec- tion by disturbing or detaching biofilms that adhere to canal surfaces and by removing a layer of infected dentin. Anatomical complexities often represent physi- cal constraints that pose a serious challenge to adequate root canal disinfection. Even with the use of rotary instrumentation, the nickel–titanium (NiTi) instru- ments currently available only act on the central body of the canal, leaving canal fins, isthmi, and cul-de-sacs untouched after completion of the preparation (7–10). These areas may harbor tissue debris and microbes and their by-products (11–13), which can prevent close adaptation of the obturation material (14) and result in persistent periradicular inflammation (15,16). In addition to mechanical preparation, irrigating solutions with a strong antibacterial effect are neces- sary. However, the available irrigants also face great challenges in eliminating all of the biofilm from the root canal. The protective mechanisms underlying biofilm antimicrobial resistance are not fully under- stood, although several mechanisms have been pro- posed (17,18). These mechanisms include physical or chemical diffusion barriers to antimicrobial penetra- tion into the biofilm, slow growth of the biofilm due to the limitation of nutrients, activation of the general stress response, and the emergence of a biofilm-specific phenotype (19). Furthermore, recent studies (20–22) have given valuable information about the interaction of endodontic disinfecting agents with dentin and other compounds present in the necrotic root canal. As a result of such interactions, the antimicrobial effec- tiveness of several key disinfectants may be weakened or even eliminated under certain circumstances. It is likely that inactivation of the medicaments in the chemical environment of the necrotic root canal is one reason for the failure to completely eradicate the microbes (23). Therefore, resistant microbes can survive on the walls of the main root canal after vig- orous chemomechanical treatment. Endodontic Topics 2012, 22, 79–98 All rights reserved 2012 © John Wiley & Sons A/S ENDODONTIC TOPICS 2012 1601-1538 79

Upload: jonathan-bravo

Post on 15-Nov-2015

52 views

Category:

Documents


4 download

DESCRIPTION

Current therapeutic options for endodontic biofilms

TRANSCRIPT

  • Current therapeutic options forendodontic biofilmsMARKUS HAAPASALO & YA SHEN

    Microbial biofilms in the infected root canal space are the primary cause of apical periodontitis. Root canal treatmenttherefore aims to either remove the biofilms from the root canal or kill all of the microbes in the biofilms.Instrumentation mechanically removes or disrupts biofilm organization and creates sufficient space in the canal toallow effective irrigation and disinfection to occur. While none of the mechanical or chemical factors alone canpredictably eradicate the infective agents, their combined action under optimal efforts is the key factor for long-termsuccess of endodontic treatment and healing of the lesion. In this article, the role and impact of various mechanical(physical) and chemical means of attacking root canal biofilms are discussed in light of relevant literature (Fig. 1).

    Received 14 November 2011; accepted 30 June 2012.

    Building blocks for success inendodontic treatment

    Apical periodontitis (AP) is an inflammatory reactionof periradicular tissues caused by a microbial infectionin the root canal (1,2). Because the bacteria in thenecrotic root canal system grow mostly in sessile bio-films, the success of endodontic treatment depends oneffective elimination of such biofilms. The necessaryelements in the control of endodontic infection arehost defense, instrumentation and irrigation, locallyused intracanal medicaments between appointments,root canal filling, and coronal restoration (3,4).Chemomechanical preparation has been regarded as

    the key element of endodontic treatment (5,6). Impor-tantly, mechanical canal preparation supports disinfec-tion by disturbing or detaching biofilms that adhere tocanal surfaces and by removing a layer of infecteddentin. Anatomical complexities often represent physi-cal constraints that pose a serious challenge to adequateroot canal disinfection. Even with the use of rotaryinstrumentation, the nickeltitanium (NiTi) instru-ments currently available only act on the central body ofthe canal, leaving canal fins, isthmi, and cul-de-sacsuntouched after completion of the preparation (710).These areas may harbor tissue debris and microbes andtheir by-products (1113), which can prevent close

    adaptation of the obturation material (14) and result inpersistent periradicular inflammation (15,16).In addition to mechanical preparation, irrigating

    solutions with a strong antibacterial effect are neces-sary. However, the available irrigants also face greatchallenges in eliminating all of the biofilm from theroot canal. The protective mechanisms underlyingbiofilm antimicrobial resistance are not fully under-stood, although several mechanisms have been pro-posed (17,18). These mechanisms include physical orchemical diffusion barriers to antimicrobial penetra-tion into the biofilm, slow growth of the biofilm dueto the limitation of nutrients, activation of the generalstress response, and the emergence of a biofilm-specificphenotype (19). Furthermore, recent studies (2022)have given valuable information about the interactionof endodontic disinfecting agents with dentin andother compounds present in the necrotic root canal.As a result of such interactions, the antimicrobial effec-tiveness of several key disinfectants may be weakenedor even eliminated under certain circumstances. It islikely that inactivation of the medicaments in thechemical environment of the necrotic root canal is onereason for the failure to completely eradicate themicrobes (23). Therefore, resistant microbes cansurvive on the walls of the main root canal after vig-orous chemomechanical treatment.

    bs_bs_banner

    Endodontic Topics 2012, 22, 7998All rights reserved

    2012 John Wiley & Sons A/S

    ENDODONTIC TOPICS 20121601-1538

    79

  • Given the inability of metal instruments to directlyplane the walls of the complex internal surface geom-etry of teeth, the key issue is the ability of antibacterialfluids to reach these spaces and surfaces to effectivelyattack bacterial biofilm. Accessory (lateral) canalsbranch from the main root canal, with diametersranging from a maximum of 100 mm to a commonminimum of 10 mm in permanent molars (24). Suchnarrow orifices create a surface tension barrier thatdoes not allow adequate mixing between the irrigantand the liquid within the canal. The narrowing of theroot canal apically (toward the root) poses a similarbarrier. Any fluid flowing down the accessory canalsfrom the root canal will have a laminar flow; turbulentflow will be not be achievable due to the very lowReynolds numbers inherent at such small pipe diam-eters, where edge effects and viscosity become themajor factors affecting fluid dynamics (25). At thescale of the accessory canals, diffusion of irrigant downthe concentration gradient will be the dominantmechanism by which the agent moves along the canal.Therefore, progress in the search for safe and moreeffective irrigant delivery and agitation systems forroot canal irrigation is necessary.In conclusion, the factors that remain a challenge in

    cleaning and disinfecting the root canal space includebiofilm resistance (17,18), poor penetration of themedicament/irrigant (26), low concentration (27),short exposure time (28,29), small overall volume

    (30), and poor exchange of irrigants in the apical partsof the root canal (24,25). In addition, inactivation ofthe medicament in the root canal in the in vivo situa-tion may weaken the effectiveness of endodontic treat-ment procedures and thus contribute to the survival ofresistant bacteria and yeasts in the root canal system.

    Locations and characteristics ofendodontic biofilmsThe localization of bacteria in the necrotic root canal isdependent mainly on ecological factors such as avail-ability of nutrients in the various parts of the canalsystem, redox potential (oxygen), and composition ofthe infective microflora including positive and negativebacterial interactions. In primary AP, the vast majorityof the microbes are occupying the main root canal(s)and only a few have invaded deeper into the dentin vialateral canals and dentinal tubules (31,32). Apicalramifications, lateral canals, and isthmuses connectingmain root canals have all been shown to harbor bac-terial cells, which are also frequently organized inbiofilm-like structures (13,33,34). In post-treatmentendodontic disease (root-filled teeth with AP), thelocation of the microbes is affected by several addi-tional factors such as the quality of the root filling,main source of the nutrients (e.g. coronal versus apicalleakage), and possible antibacterial effects of the rootfilling materials (16). In addition, biofilms attached tothe apical root surface (extraradicular biofilms) havebeen reported and regarded as a possible cause ofpost-treatment apical periodontitis (35). However, it isimportant to understand that in post-treatment endo-dontic disease, the bacteria (or yeasts) must be able tointeract with the host defense of the paradental tissuesto cause the inflammatory response and tissue destruc-tion, which can then be detected in the radiograph.A necrotic root canal also represents a challenging

    environment in which bacteria face toxic substancessuch as bacteriocins and have to survive with limitedaccess to nutrients and key elements such as iron. Thisis likely to result in various survival strategies such asdecreasing metabolic activity and even transforminginto the viable but non-culturable state (36). Chvezde Paz et al. (37) also found that the low reactivity ofnon-growing biofilm cells to the introduction of freshnutrients may be a survival strategy employed bymicroorganisms in the oral cavity. These various phasesof microbial interaction with the surface appear to

    Fig. 1. Green symbols: factors that effect the develop-ment and characteristics of endodontic biofilms. Redsymbols: factors that are important in eradication ofbiofilms and biofilm bacteria.

    Haapasalo & Shen

    80

  • require the production of extracellular polymers thatassist in initial adhesion, maintenance of biofilm struc-ture, and detachment from matrix-enclosed aggre-gates. Overall, the unique root canal environmentalconditions are expected to influence biofilm structureand function (16). Endodontic biofilm morphologydiffers considerably from individual to individual, andthe reasons for that deserve further investigation butmay conceivably be related to different biofilm com-position, type and availability of nutrients, and overallduration of the infection (16,26,38).

    Effect of instrumentationon biofilmsThe purpose of root canal preparation in the contextof endodontic therapy is to: (i) shape the canals to anadequate geometry; (ii) clean the canal system by pro-moting access for disinfection solutions (this strategyhas been termed chemomechanical canal preparation);and (iii) make it possible to place a high-quality rootfilling (4). Microbiologically, the goal of instrumenta-tion and irrigation is to remove or kill all of the micro-organisms in the root canal system, and neutralize anyantigenic/biological potential of the microbial com-ponents remaining in the canal. If this goal couldpredictably be achieved at the first appointment, mosttreatments could be completed in one visit, if only thetime available would allow it. In cases where this (com-plete eradication of root canal microorganisms) cannotbe achieved, the goal of instrumentation and irrigationis to create optimal conditions for the placement of anantibacterial inter-appointment dressing in order tofurther enhance the disinfection of the canal.Mechanical instrumentation is the core method for

    bacterial reduction in the infected root canal. With thelaunch of nickeltitanium (NiTi) rotary systems,perhaps too much credit was given to these systems asbeing the sole solution to challenges in root canaltreatment. Regarding the direct efficacy in the removalof bacteria, it is important to notice that no differencewas found between hand and rotary instruments (39).Dalton et al. (40) compared the ability of stainless-steel K-type files and NiTi rotary instruments toremove bacteria from infected root canals using salineas the irrigating solution. The canals were sampled formicrobes before, during, and after instrumentation. Inthis study, only about one-third of the canals wererendered bacteria-free, and no significant difference

    was detected between canals instrumented with handfiles and rotary instruments. An in vivo study thatapplied correlative light and electron microscopictechniques to evaluate residual intracanal infectionafter instrumentation with stainless-steel hand files inmesio-buccal canals and NiTi instruments in mesio-lingual canals of the same lower molars showed thatthere was no difference in their respective ability toeliminate infection (33). In addition, Carver et al. (41)evaluated the in vivo antibacterial efficacy of a hand/rotary technique in mesial root canals of necrotic man-dibular molars. Root canal cleaning and shaping withhand and rotary instrumentation and irrigation with6.0% sodium hypochlorite showed a significant reduc-tion in the log colony-forming unit (CFU) counts.However, bacteria still remained in the canals.Moreover, mechanical disinfection can also be

    related to the removal of a layer of infected dentin, orat least of incompletely mineralized predentin (6). Ithas been shown that bacteria might penetrate dentinaltubules to depths of 200 mm or more (42,43). Com-plete uniform enlargement of a root canal by 200 mmis not achieved with any contemporary instrument;this appears to be an unattainable goal for anymechanical canal preparation technique (44,45).It has been shown that the amount of mechanically

    prepared canal surface and, perhaps equally, theamount of disturbed biofilm in the main root canaldepends on the canal type (10,45). Rotary instru-ments perform comparably poorly in long oval canalssuch as distal canals in lower molars, specificallybecause they do not mechanically prepare 60% or moreof the canal surface under these conditions (45,46). Inthe case of an infected root canal, any bacterial biofilmon the instrumented canal surfaces is likely to be dis-turbed or removed, although some of the bacterialcells may become embedded within the smear of tissue(47). The bacterial biofilm on the uninstrumentedsurfaces is likely to remain mechanically undisturbed,except by the displacement of any pulpal tissue ordentinal debris from the prepared part of the canal. Itis possible that changes in the ecology of the root canalsystem may have a significant influence on the survivaland death of bacteria on the uninstrumented surface.Nevertheless, the uninstrumented surfaces should stillbe regarded as contaminated.A newly developed self-adjusting file (SAF)

    (ReDent-Nova, Raanana, Israel) was designed toaddress the shortcomings of traditional rotary files by

    Current therapeutic options for endodontic biofilms

    81

  • adjusting itself to the cross-section of the canal (48).The SAF system uses a hollow vibrating instrument,which allows for continuous irrigation with NaOCl orethylenediaminetetraacetic acid (EDTA) throughoutthe instrumentation process (Fig. 2). Irrigants areexchanged and taken to the apical root canal as a resultof the vibration and in-and-out motion of the SAF.The compressible NiTi tube can adapt itself to theoval-shaped canal while its abrasive blades are pressedagainst the walls to promote root canal enlargement.When compared with NiTi instrumentation, it hasbeen reported that the SAF leaves fewer unpreparedareas in anterior teeth (49) and molar root canals(48,50). Siqueira et al. (51) compared the capability ofSAF and rotary NiTi instrumentation to eliminateEnterococcus faecalis populations from extractedhuman teeth. Long oval canals from mandibular inci-sors and maxillary second premolars were infectedwith E. faecalis for 30 days in order to form biofilm-like structures. Preparation of long oval canals withthe SAF was significantly more effective than rotaryNiTi instrumentation in reducing intracanal E. faecaliscounts. Data regarding the incidence of negativeand positive cultures revealed that in the SAF group,80% of the samples were rendered free of detectablelevels of E. faecalis, whereas instrumentation withrotary NiTi instruments resulted in only 45% of thesamples being culture-negative. The SAF system hasthe potential to be particularly advantageous in pro-moting the disinfection of oval-shaped canals.

    However, it is presently unknown whether canalpreparation with the SAF, and in particular its poten-tial to debride canal walls better, will lead to improvedclinical outcomes.The reported incidence of isthmuses in the mesial

    root of mandibular molars ranges from 5489%,mostly in the middle and apical thirds (52). In addi-tion to these structures being inaccessible to instru-ments, instrumentation can in fact further complicatethe cleaning of these areas. Paqu et al. (45) showedthat dentin debris is formed and packed into theisthmus area during rotary instrumentation withoutirrigation. Accumulated debris certainly has a negativeimpact on the sealability of root canals, but it may alsohamper disinfection in cases with apical periodontitis.Endal et al. (53) found that even copious irrigationduring and after instrumentation with solutions dis-solving both organic and inorganic matter was not ableto prevent or remove the debris packed into theisthmus area between the main root canals (Fig. 3),where bacteria may be present in the form of biofilms(33). Although mechanical instrumentation togetherwith the use of irrigants in the canal is often quiteeffective, complete cleanliness of these inaccessibleareas is difficult to achieve. In an in vivo study, Burle-son et al. (54) examined the efficiency of hand/rotarytechniques in removing biofilm/necrotic tissue in themesial roots of necrotic human mandibular molars.Following extraction, histological preparation, andstaining, cross-sections from the 1- to 3-mm apicallevels were evaluated for percentage of biofilm/necrotic debris removal. Cleanliness results at the 1-,2-, and 3-mm levels were 80%, 92%, and 95% forcanals and 33%, 31%, and 45% for isthmuses, respec-tively. Interestingly, a recent case report (55) showedthat a complex, variable, multi-species biofilm waspresent along the entire length of the isthmus in whichthe tooth had been initially treated 10 years earlier andthen re-treated 2 years ago. Both Gram-positive andGram-negative organisms were detected, surviving inan extremely harsh and nutrient-deficient environmentthat had existed for more than a decade after root canaltreatment.In summary, instrumentation plays an important

    role in helping to remove biofilm from those areaswhere the instrument can gain direct contact with theroot canal wall. In addition, shaping of the main canalfacilitates effective irrigation by creating the necessaryspace for needle penetration and sufficient irrigant

    Fig. 2. Self-adjusting file system. The hollow structureof the file and elasticity of NiTi allow the file to comeinto contact with the root canal dentin walls in a largerarea than is possible with traditional rotary files.

    Haapasalo & Shen

    82

  • flow. Regardless, challenges remain in many areas dueto anatomy and the resistance of biofilms.

    Effect of various irrigatingsolutions on root canal biofilmsAntimicrobial agents have often been developed andoptimized for their activity against fast-growing, dis-persed populations of a single species. However,microbial communities growing in biofilms areremarkably more difficult to eradicate with antimicro-bial agents, and microorganisms in mature biofilmscan be extremely resistant for reasons that have yet tobe fully explained. Most of the endodontic studies onbiofilms have been conducted with monocultures byallowing cells to grow on membranes, glass, or plasticand divide under a continuous or frequent supply offresh nutrients from a few hours to a few days(27,28,5659). As the influence of root canal dentinand other surfaces on the expression of novel biofilmphenotypes has not yet been touched upon, conclu-sions and decisions reached from studies of monocul-ture biofilm in the laboratory must be taken with greatcaution. Such models may not reflect the reality of theinfected root canal and thus may give misleading inter-pretations, especially regarding the effects of antimi-crobials on biofilm bacteria. Therefore, it is importantto develop multi-species in vitro biofilm models with aclose similarity to oral/endodontic in vivo biofilms(Fig. 4).

    Given that the current instrumentation techniquesalone are unable to render root canals bacteria-free, achemical irrigant is regarded as necessary to assist inreducing the amount of bacteria and their toxicby-products. In addition, an ideal irrigant shouldremove organic and inorganic debris and have low orno tissue toxicity (60). While none of the irrigatingsolutions/disinfecting agents presently used in endo-dontic treatment are able to do all of the requiredtasks, many of them can have an impact on biofilms,either by dissolving the film, killing the microbes resid-ing in the biofilm, or by helping to break down ordetach the film from the surface.

    Sodium hypochlorite

    Sodium hypochlorite (NaOCl) is the most popularand important irrigating solution (61). In water,NaOCl ionizes into the sodium ion, Na+, and thehypochlorite ion, OCl-, establishing equilibrium withhypochlorous acid (HOCl). Hypochlorous acid isresponsible for the antibacterial activity; the OCl- isless effective than undissolved HOCl.NaOCl is commonly used in concentrations between

    0.5% and 6%. It is the only irrigant in Endodontics thatcan dissolve organic tissue, including the organic partof the smear layer. It should be used throughout theinstrumentation phase. Dunavant et al. (27) comparedthe efficacy of 1% or 6% NaOCl with that of 2% chlor-hexidine (CHX), Smear Clear, and MTAD against

    Fig. 3. Three-dimensional reconstruction micro-CT scans of the mesial root canal system of the mandibular molarunder investigation. (A) Three-dimensional micro-CT reconstruction after instrumentation. Prepared canal areas areindicated in blue, and untouched areas are indicated in red. (B) Superimposition of the apparent accumulated hardtissue debris areas are indicated in yellow. The canal space and empty space in the isthmus area after instrumentationare indicated in silver. (C) Root filling material is indicated in silver; the non-filled area in the ribbon-shaped isthmusthat includes debris and void is shown in dark violet.

    Current therapeutic options for endodontic biofilms

    83

  • E. faecalis biofilms in an in vitro model system. Theirmodel consisted of biofilms grown in a flow cellsystem. Biofilms were immersed in test irrigants for 1or 5 minutes. Results showed that both concentrationsof NaOCl provided statistically significantly betterbiofilm killing than any of the other agents tested. 6%NaOCl also removed biofilm cells. In an ex vivobiofilm study, Clegg et al. (62) demonstrated a differ-ence in the effectiveness of 6% and 3% NaOCl againstbiofilm bacteria, the higher concentration being moreeffective. Bacterial cells in the root canal encounter aharsh ecological milieu. Recently, one in vitro study(38) evaluated the biofilm formation capability ofstarved E. faecalis cells on human dentin and the sus-ceptibility of the biofilm to 5.25% NaOCl. The find-ings showed that E. faecalis cells in the starvationphase could develop biofilm on human dentin. Bio-films of starved cells were more resistant to 5.25%NaOCl than those of stationary cells.Despite some good in vitro results, the more limited

    antimicrobial effectiveness of NaOCl in vivo is disap-pointing. The poor in vivo performance compared tothe in vitro effect may be caused by problems in pen-etration to the most peripheral parts of the root canalsystem such as fins, anastomoses, apical canals, lateralcanals, and dentin canals. Also, the presence of inacti-vating substances such as exudate from the periapicalarea, pulp tissue, dentin collagen, and microbialbiomass counteracts the effectiveness of NaOCl (63).One of the unknown areas regarding the effect of

    NaOCl on biofilms is the role of EPS (extracellularpolymeric substance) in this interaction. Althoughsome in vitro studies have shown a complete disap-pearance of the biofilm with strong sodium hypochlo-rite treatment, one cannot exclude the possibility thatwhile some/most of the biofilm may have been dis-solved, the effect can also be caused by detachment ofthe biofilm from its substrate in the in vitro environ-ment. Nevertheless, as NaOCl is the only solution inEndodontics that can at least to some extent dissolvebiofilm in addition to direct killing of microbes insidethe film, it should be regarded as the main disinfectingsolution during chemomechanical preparation ofinfected root canals.

    Chlorhexidine digluconate and CHX-Plus

    Chlorhexidine digluconate is widely used in disinfec-tion in Dentistry because of its antimicrobial activity(6466). It has gained considerable popularity inEndodontics as an irrigating solution and as an intra-canal medicament. However, CHX has no tissue-dissolving capability and therefore it cannot replacesodium hypochlorite.CHX permeates the microbial cell wall or outer

    membrane and attacks the bacterial cytoplasmic orinner membrane or the yeast plasma membrane. Inhigh concentrations, CHX causes coagulation of intra-cellular components (67). One of the reasons for thepopularity of CHX is its substantivity (i.e. continued

    Fig. 4. Scanning electron micrographs of biofilms with mixed bacterial flora including numerous spirochetes.(A) Three-week-old biofilm. (B) Six-week-old biofilm after 2% CHX treatment for 3 minutes shows tightly coiledspirochetes and a few damaged bacterial cells.

    Haapasalo & Shen

    84

  • antimicrobial effect) because CHX binds to hard tissueand remains antimicrobial. However, similar to otherendodontic disinfecting agents, the activity of CHXdepends on the pH and is also greatly reduced in thepresence of organic matter (65). Several studies havecompared the antibacterial effect of NaOCl and 2%CHX against intracanal infections and have shownlittle or no difference between their antimicrobialeffectiveness (6871). Clegg et al. (62) evaluated theex vivo effectiveness of sodium hypochlorite, CHX,and MTAD against biofilms grown on apical dentin.Six percent sodium hypochlorite was the only solutioncapable of disrupting and completely removing thebiofilm after 15 minutes of exposure. 2% CHX killedthe biofilm bacteria but was not able to disrupt thebiofilm structure (62). Although CHX may kill thebacteria, the biofilm and other organic debris are notremoved by it. Residual organic tissue may have anegative effect on the quality of the permanent rootfilling seal, necessitating the use of NaOCl duringinstrumentation.Surface-active agents have been added to several dif-

    ferent types of irrigants in order to lower their surfacetension and to improve their penetration into the rootcanal. Recently, a few studies have been published inwhich the antibacterial activity of a chlorhexidineproduct with surface-active agents (CHX-Plus, VistaDental Products, Racine, WI) has been compared toregular CHX, both with a 2% chlorhexidine concen-tration. One study (29) showed superior killing ofbiofilm bacteria by the combination product. Anotherstudy (26) examined the susceptibility of multi-species

    biofilm bacteria at different phases of biofilm growthto 2% CHX and CHX-Plus. The multi-species biofilmswere grown from plaque bacteria on collagen-coatedhydroxyapatite discs in brainheart infusion broth fortime periods ranging from 2 days to several months.Fresh nutrients were added weekly for the first 3weeks, followed by a nutrient-deprivation phase, whenfresh medium was added only once a month. Biofilmsof different ages were subjected to a 1-, 3-, or10-minute exposure to 2% CHX or CHX-Plus. Theproportion of killed bacteria in mature biofilms (3weeks or older) was lower than in young biofilms (2days, 1 or 2 weeks) after treatment with both CHXproducts, the reduction being much greater with theregular 2% CHX (26). The resistance of mature bio-films under the nutrient-limiting phase (612 weeks)to CHX remained stable and was similar to 3-week-oldbiofilm (Fig. 5). CHX-Plus showed higher levels ofbactericidal activity at all exposure times compared to2% CHX, which may indicate that the surfactant com-ponent in CHX-Plus facilitated penetration of thedisinfectant into the biofilm. Overall, this study dem-onstrated that bacteria in mature biofilms andnutrient-limited biofilms are more resistant to CHXkilling than bacteria in young biofilms. The result alsoemphasizes the importance of standardizing the age ofthe biofilm cultures to allow comparisons betweenstudies. It is likely that the biofilms in in vivo rootcanals are almost always older than 3 weeks; thereforethe results of in vitro experiments with biofilmsyounger than 3 weeks should be evaluated withcaution.

    Fig. 5. The proportion of viable cells (volume) of biofilms of different ages after treatment with CHX and CHX-Plus.There was a significant difference in the proportion of killed bacteria depending on the type of the disinfecting agent,exposure time, and age and nutrient supply of the biofilm.

    Current therapeutic options for endodontic biofilms

    85

  • MTAD

    Bio Pure MTAD (Dentsply Tulsa Dental, Tulsa, OK)was introduced to Endodontics in 2003. This mixtureof a tetracycline isomer (doxycycline), citric acid, and adetergent has been shown to be effective in smear layerremoval (72,73). Some in vitro experiments indicatedthat MTAD has a strong antibacterial effect. It hasbeen suggested to be more effective than NaOCl andEDTA against E. faecalis (74) and mixed bacteria (75).However, some of these results were later challengedin studies which found the antibacterial effect ofMTAD to be inferior to 6% NaOCl and 2% chlorhexi-dine (27). Furthermore, Dunavant et al. (27) reportedthat 1% NaOCl killed six times more E. faecalis inbiofilms (99.78%) than MTAD did (16.08%). Pappenet al. (76) tested the efficacy of MTAD, Tetraclean,and five experimental solutions against biofilm bacte-ria. Tetraclean was more effective against two-week-old polymicrobial biofilm than MTAD. A comparisonof biofilm killing by MTAD and the experimental solu-tions indicated that the type of detergent in the medic-ament mixture may have been of major importance inthe effectiveness of the solutions against biofilms (76).

    QMiX

    QMiX (Dentsply Tulsa Dental) is a new irrigatingsolution which contains EDTA, chlorhexidine, and adetergent (surface-active agent); its pH is slightlyabove neutral (Fig. 6) (77,78). A surface-active agentdecreases the surface tension of solutions and increasestheir wettability (79). Also, it enables better penetra-tion of an irrigant into the root canal (80).The effect of QMiX against E. faecalis and mixed

    plaque biofilms was evaluated in a recent study usingthree-week-old biofilms grown on collagen-coatedhydroxyapatite discs under anaerobic conditions (80).The killing of bacteria inside the biofilm was measuredby confocal laser scanning microscopy (CLSM) andviability staining. The results demonstrated that QMiXand 2% NaOCl were superior to 1% NaOCl, 2% CHX,or MTAD by killing two to twelve times more biofilmbacteria in one to three minutes. 2% NaOCl was moreeffective than QMiX at 1 minute against plaquebiofilm bacteria, but at 3 minutes QMiX had killedmore bacteria (65.3%) than any other solution tested(80). NaOCl solutions stronger than 2% could not betested in the model because of the strong bubbleformation, which made CLSM impossible.

    The presence of bacteria in the dentinal tubules hasbeen associated with persistent root canal infection(81). Studies have shown that bacteria can penetrateinto dentinal tubules, and the depth of penetrationvaries from 200 mm to 1,500 mm (42,43,82). Bacteriawithin the dentinal tubules may be poorly accessible toroot canal irrigants, medicaments, and sealers becausethey may have limited penetrability into the dentinaltubules. An in vitro study (77) using a novel type ofdentin infection model found that QMiX was equallyeffective at killing E. faecalis bacteria in dentin as 6%NaOCl: over 40% and 60% of the bacteria were killedby both at 1 minute and 3 minutes, respectively(Fig. 7). Both solutions were more effective againstthe bacteria inside dentin than 1% or 2% NaOCl or 2%CHX.

    Ethylenediaminetetraacetic acid (EDTA)

    EDTA is a calcium binder (chelator) that aids in theremoval of the smear layer. The smear layer is mainlycomposed of dentin particles embedded in an amor-phous mass of organic material that forms on the innerroot canal walls during the instrumentation procedure.The smear layer counteracts disinfectants and mayblock or slow down the penetration of medicamentsinto the dentinal tubules (42). It also interferes withthe adhesion of some and penetration of all root filling

    Fig. 6. Two combination products: (A) CHX-Plus; and(B) QMiX.

    Haapasalo & Shen

    86

  • materials. Therefore, by facilitating the cleaning andremoval of infected tissue, EDTA contributes to theelimination of bacteria in the root canal. It has alsobeen shown that the removal of the smear layer byEDTA improves the antibacterial effect of locally useddisinfecting agents in deeper layers of dentin (42,83).Chemicals that alter the physicochemical properties

    of dentin might influence the nature of the bacterialadherence and the adhesion force to dentin, which arefactors in biofilm formation. Kishen et al. (84) inves-tigated the effects of endodontic irrigants on theadherence of E. faecalis to dentin. The bacteria adher-ence assay was conducted by using fluorescencemicroscopy, and the adhesion force was measured byusing atomic force microscopy. There were significantincreases in adherence and the adhesion force afterirrigation of dentin with EDTA, whereas NaOClreduced them. With the use of CHX, the force ofadhesion increased, but the adherence assay showed areduction in the number of adhering bacteria.However, the sequence in which NaOCl and EDTAare used for canal irrigation has an impact on the levelof dentin erosion on the main root canal wall (85).Sodium hypochlorite used as a final irrigant solutionafter demineralization agents causes marked erosion of

    root canal dentin. So far, it is not known whether sucherosion is harmful to the root dentin and the tooth.However, chemical removal, even by strong surfaceerosion, may facilitate the removal of biofilms from theuninstrumented parts of the root canal.

    Mechanical agitation by sonic andultrasonic appliancesThe hydrodynamic behavior of the irrigating solutionsplays an important role in the effectiveness of irrigation(15,25). It also depends on the working mechanism ofthe irrigant as well as the mechanism of action of theequipment used to introduce and agitate the irrigantin the canal (86). Irrigant agitation can be done manu-ally with a needle and syringe or by machine-drivenforces such as in sonic and ultrasonic agitation.

    Sonic agitation

    In 1985, Tronstad et al. (87) were the first to reporton the use of a sonic instrument in Endodontics. Sonicirrigation is different from ultrasonic irrigation in thatit operates at a lower frequency (110 kHz) and pro-duces smaller shear stresses (88). The sonic energy alsogenerates significantly higher amplitude or greaterback-and-forth tip movement.Sonic activation has been shown to be an effective

    method for disinfecting root canals (89). The Endo-Activator (Advanced Endodontics, Santa Barbara, CA)uses sonic energy to agitate the irrigants in the rootcanal system (Fig. 8). The action of the EndoActivatortip often produces a cloud of debris originating fromthe canal contents. Vibrating the tip, in combinationwith moving the tip up and down in short verticalstrokes, synergistically produces a powerful hydrody-namic phenomenon (30). It has been suggested that10,000 cycles per minute is needed to optimize thedebridement and promote disruption of the smearlayer (90). The EndoActivator System has beenreported to be able to clean debris from lateral canals,remove the smear layer, and dislodge clumps of simu-lated biofilm within the curved canals of molar teeth(90). However, not all studies have reported similarresults. Brito et al. (91) found that sonic activation ofEDTA and NaOCl with the EndoActivator deviceafter chemomechanical procedures on a straight singlecanal did not result in improved disinfection comparedto conventional needle irrigation. An in vivo study

    Fig. 7. Viability staining and confocal laser scanningmicroscopy of E. faecalis infected dentinal tubulestreated with different antibacterial solutions for 3minutes each: (A) sterile water; (B) 2% NaOCl; (C) 6%NaOCl; and (D) QMiX.

    Current therapeutic options for endodontic biofilms

    87

  • (92) reported that the EndoActivator did not enhancethe ability of standard needle irrigation to eliminatecultivable bacteria from root canals.Shen et al. (93) investigated whether mechanical agi-

    tation (ultrasonic or sonic) improved the effectivenessof chlorhexidine against biofilm bacteria in vitro. Formechanical agitation, an ultrasonic tip or an EndoAc-tivator (sonic) tip was placed 5 mm above the top ofthe multi-species biofilm, which was immersed in irri-gant. This was the minimum distance between theultrasonic or sonic tip and the biofilm surface at whichmechanical agitations did not disrupt or disperse thebacteria. After treatment, the amount of dead bacteriain biofilms was analyzed by viability staining and con-focal laser scanning microscopy. The low-intensityultrasonic or sonic agitation that does not disrupt ordisperse the biofilm bacteria improves the action ofdisinfectants against biofilm bacteria. The precisemechanisms of the enhanced killing have not beenidentified and may be different in different situations.When the EndoActivator tip was placed 5 mm over thetop of the biofilm in this study, the acoustic stream wasconnected to the rapid movement of the irrigatingsolution in a vortex around the biofilm (93). Thisenhanced transport may be partially responsible for theincreased killing of biofilm bacteria exposed to com-binations of the disinfectant and mechanical agitation.The visual observation of more bubbles exiting alongthe EndoActivator file during irrigation indicates thatbubbles do not exit in a perfect linear stream but theirflow is turbulent and chaotic; thus, creating a columnof bubbles instead of a line of bubbles produces abetter result. The formation of micro-bubbles gradu-

    ally increasing in diameter until they collapse provokesvery effective small implosions, which produce anirregular agitation of the irrigant. The results showedthat the combined use of ultrasonic or sonic vibrationand chlorhexidine produced a better antimicrobialeffect against biofilms than chlorhexidine alone. Thevolume of killed cells was significantly correlated withthe time of exposure, the type of medicament, and thetreatment group (sonic, ultrasonic, or no mechanicalagitation) (93).

    Ultrasonic agitation

    Ultrasonics in Endodontics was introduced byRichman in 1956 for the preparation of the accesscavity and for preparation as well as obturation of thecanals (94). Twenty years later, Martin (95) describedthe in vitro disinfectant action of ultrasonics, demon-strating that the combined use of ultrasonics andsodium hypochlorite could be more effective thaneither one on its own.The ultrasonics used in Endodontics are acoustic

    vibrations with frequencies around 25,000 cycles/second. From the energy source, the ultrasonic wavesare transferred via a transducer to a liquid, wherebywell-known physical phenomena occur. One of these isan acoustic stream and is connected to the rapidmovement of fluid particles in a vortex around theobject that vibrates (88). Another phenomenoncaused by the ultrasonic vibration is cavitation, whichis the formation of micro-bubbles that graduallyincrease in diameter until they collapse, provokingvery effective small implosions that produce an irregu-

    Fig. 8. (A) The EndoActivator system has different-sized tips and one handpiece. (B) EndoActivator with a smallplastic tip.

    Haapasalo & Shen

    88

  • lar agitation of the liquid. Both of these effects areindicated (88,95) as the principal reason why thedebris are removed from the dentinal walls. It shouldalso be remembered that ultrasonics raise the tem-perature of the liquid that surrounds the vibratingobject (96).Numerous investigations have demonstrated that

    the use of passive ultrasonic irrigation (PUI) after handor rotary instrumentation resulted in a significantreduction in the number of bacteria (96100) orachieved significantly better results than syringe needleirrigation (99101). Carver et al. (41) found in vivothat the use of ultrasonic irrigation following hand/rotary instrumentation produced a significantlygreater reduction in CFU counts in infected necrotichuman molars. Additionally, a significantly higher per-centage of canals showed no microbial growth insamples taken from the canals following the additionof ultrasonic irrigation (80%) than following hand/rotary instrumentation alone (27%). Histologicalspecimens from an in vivo study by Burleson et al. (54)confirmed that one-minute use of ultrasonically acti-vated irrigation following hand/rotary root canalcleaning and shaping improved canal and isthmuscleanliness because less necrotic debris/biofilm wasleft behind.The most effective mechanism behind the effect of

    ultrasound is the formation of cavitation bubbles(102104). These bubbles are in a non-equilibriumstate and will oscillate and collapse. The bubbledynamics involved is often complex because of theproximity of the nearby tissues (105). The forcefulbubble collapse with high-speed jetting could be har-nessed beneficially (as in the removal of biofilm) orcould cause undesirable collateral damage (106).High-intensity focused ultrasound (HIFU) is appliedclinically to generate collapsing cavitation bubbles influids and tissues, which collapse with high-speed jetsthat can be used for drug delivery. However, itshould be remembered that the distance from theultrasonic tip where cavitation can occur is veryshort, in the range of 10100 mm (107). Recently,Shrestha et al. (108) found that the collapsing cavi-tation bubbles used in HIFU treatment resulted insignificant penetration (up to 1,000 mm) of antibac-terial nanoparticles into the dentinal tubules. Thefindings demonstrated the potential application ofHIFU-generated collapsing cavitation bubbles todeliver antibacterial nanoparticles into the dentinal

    tubules and subsequently improve disinfection inEndodontics.

    Photo-activated disinfectionPhoto-activated disinfection (PAD) involves the use ofa photo-active dye (photosensitizer) that is activatedby exposure to light of a specific wavelength in thepresence of oxygen. The transfer of energy from theactivated photosensitizer to available oxygen results inthe formation of toxic oxygen species, such as singletoxygen and free radicals. These very reactive chemicalspecies can damage proteins, lipids, nucleic acids, andother cellular components (109111). PAD has beenintroduced to Dentistry as a host-friendly way ofattacking microorganisms in periodontal and endo-dontic infections. While most other substances ormethods used in root canal disinfection are directly orpotentially harmful to the host, PAD is claimed tospecifically target microorganisms with no collateraldamage. It involves the use of a photosensitizer (PS)that is activated by light in the presence of oxygen.There are several factors influencing photodamage

    including the type, dose, incubation time, and local-ization of the photosensitizer; the availability ofoxygen; the wavelength of light; the light powerdensity; and the light energy fluence. An importantcharacteristic of photodynamic therapy is its inherentdual selectivity; first, by achieving an increased con-centration of the photosensitizer by specific binding totarget tissues, and second, by constraining the irradia-tion to a specified volume. In antibacterial photody-namic therapy, photodestruction is mainly caused bydamage to the cytoplasmic membrane and DNA(112,113).The efficiency of PAD may depend on environmen-

    tal and microbiological factors at the site of the infec-tion. A fundamental difference in susceptibility to PADbetween Gram-positive and Gram-negative bacteriahas been reported (114,115). In general, neutral,anionic, or cationic PS molecules can efficiently killGram-positive bacteria, whereas only cationic PS mol-ecules or strategies that permeabilize the Gram-negative permeability barrier in combination withnon-cationic PS molecules are able to kill multiple logsof Gram-negative species. This difference in suscepti-bility between species in the two bacterial classifica-tions was explained by their physiology, as theGram-positive species have a cytoplasmic membrane

    Current therapeutic options for endodontic biofilms

    89

  • surrounded by a relatively porous cell wall composedof peptidoglycan and lipoteichoic acid that allows PSmolecules to cross. The cell envelope of Gram-negative species is composed of an outer membrane, athin peptidoglycan layer, and a cytoplasmic membraneas the innermost cell wall structure. The movementof molecules across the Gram-negative cell wall isstrictly regulated at the outer membrane that is richin lipopolysaccharides (LPS) (116,117). Negativelycharged LPS molecules have a strong affinity forcations such as calcium (Ca2+) and magnesium (Mg2+),the binding of which is required for the thermody-namic stability of the outer membrane. Antimicrobialphotosensitizers such as porphyrins, phthalocyanines,and phenothiazines (e.g. Toluidine blue O and Meth-ylene blue), which bear a positive charge, can directlytarget both Gram-negative and Gram-positive bacteria(118,119). Toluidine blue O and Methylene blue arecommonly used for oral antimicrobial photodynamictherapy. The functioning of self-promoted up-takepathways and protein transporters is modulated bycharged entities such as cations. Therefore, the successof PAD in eliminating bacteria from anatomical sitessuch as root canals could be influenced by the cation-rich microenvironment persisting at these sites.PAD conducted on endodontic biofilms has often

    failed to achieve effective microbial killing, promptingmany researchers to combine PAD with conventionalantimicrobial strategies for superior performance(112,120123). Methylene blue has been used as thephotosensitizer for targeting endodontic micro-organisms in several studies (122,124126). The pho-todynamic effects of Methylene blue were investigatedon multi-species root canal biofilms comprised of fourspecies of microorganisms in experimentally infectedroot canals of extracted human teeth (122). PADachieved a reduction in bacterial viability of up to 80%.The results of this study suggested the potential ofPAD to be used as an adjunctive antimicrobial proce-dure after standard endodontic chemomechanicaldebridement, but also demonstrated the importanceof further optimization of light dosimetry for bacterialphotodestruction in root canals. Also, modified PSformulations with improved photochemical and pho-tobiological properties have shown that the nature ofthe PS solvent used for PAD influences its bactericidalpotential (123,126,127). Methylene blue dissolved ina mixture of glycerol, ethanol, and water (123,125), aswell as a Methylene blue formulation containing an

    emulsion of oxidizer and oxygen carrier (125),enhanced the photodynamic effects of Methylene bluein vitro. Findings from a recent study showed theefficacy of photodynamic therapy mediated byMethylene blue dissolved in a mixture of glycerol,ethanol, and water in the presence of an irradiationmedium (perfluorodecahydro-napthalene) to eradicateE. faecalis biofilms in the root canal system of experi-mentally infected human teeth (128). The use ofMethylene blue mediated PAD with modified PSformulations was found to enhance the efficacy ofPAD in destroying Gram-positive E. faecalis biofilmand Gram-negative Pseudomonas aeruginosa biofilm(115).PAD as an adjunctive technique to standard endo-

    dontic treatment may have potential in the clinicalsetting by providing a large therapeutic windowwhereby residual root canal bacteria can be killedwithout harming cells in the periapical region. Futureexperimental studies should explore the use of noveltechnologies for increased delivery of Methylene blueor Toluidine blue O in dentinal tubules and the appli-cation of supplemental hyperoxygenation in the rootcanal system to enhance the photodynamic therapyeffect.

    Local intracanal medicamentsIn the treatment of teeth with a vital pulp, there is noneed for intracanal antibacterial medication. However,in the treatment of apical periodontitis, intracanalmedication has been recommended by many in orderto eradicate the microbes that survive instrumentationand irrigation. A variety of medicaments have beenused for this purpose. These include calcium hydrox-ide [Ca(OH)2]; phenol compounds such as eugenol,camphorated parachlorophenol (CMCP), and formo-cresol; iodine potassium iodide (IPI); glutaraldehyde;and pastes containing a mixture of antibiotics with orwithout corticosteroids.As an inter-appointment dressing, a substance

    should be selected that is not easily replaced by tissuefluid and that can remain physically intact over weeksor months. A water slurry of Ca(OH)2 combinesseveral attractive features (129,130) of a good intra-canal dressing. It is strongly alkaline (pH 12.5) anddissociates into calcium and hydroxide ions in aqueoussolution; the latter provide antimicrobial effects (131)and tissue-dissolving capacity (132). With its fairly low

    Haapasalo & Shen

    90

  • solubility and mere physical presence, it may be used asan intracanal dressing over long periods of time. Itsmost essential function is then to obstruct bacterialregrowth. The antimicrobial activity of calciumhydroxide seems dependent upon direct contact withbacteria (130). Direct contact experiments in vitro(59) showed that Ca(OH)2 was 100% effective ineliminating 2-day-old E. faecalis biofilm in a mem-brane filter model. This could be attributable to thefact that the pH of Ca(OH)2 remained high in themembrane.However, owing to its poor solubility and diffusibil-

    ity, Ca(OH)2 is a rather inefficient antimicrobialagainst microorganisms lodged in pulpal remnants,crevices of the canal, and dentinal tubules (83,133). Aswell, the buffering ability of tissues impacts pH levelchanges. Using scanning electron microscopy andscanning confocal laser microscopy, Distel et al. (134)reported that despite intracanal dressing withCa(OH)2, E. faecalis formed biofilms in root canals.Both enterococci and yeasts sustain a high alkalineenvironment and are able to survive in root canalsmedicated with Ca(OH)2 (135,136). These results areamongst the reasons why controversy has emergedover its usefulness as an antimicrobial agent in rootcanal treatment. Although several clinical trials haveobserved that root canals are rendered free of culti-vable bacteria following its application for a week ormore (131,137), others have found that micro-organisms can still be recovered from a substantialnumber of medicated root canals (138140). Differ-ences in findings may relate to the type of teethincluded in the studies and the associated effectivenessof the biomechanical preparation, sampling technique,and the extent to which Ca(OH)2 was eliminated fromthe root canals prior to the sampling procedure.The clearly poorer results in vivo in the root canal

    indicate the presence of interfering factors that nega-tively affect the outcome of the disinfection. Haapasaloet al. (63) and Portenier et al. (20,21) studied theeffect of dentin and other substances present in theroot canal milieu on the antibacterial effect of com-monly used intracanal medicaments such as calciumhydroxide, chlorhexidine, and IPI against E. faecalis.These studies showed that all three disinfectants werenegatively affected by the various substances tested,calcium hydroxide being particularly sensitive to theinhibitory effect of a variety of substances present inthe root canal.

    Adherence to dentin and inter-species interactionsin a biofilm appear to differentially affect the sensitiv-ity of microbial species to calcium hydroxide. Brandleet al. (57) investigated the effects of growth condi-tions (planktonic, mono- and multi-species biofilms)on the susceptibility of E. faecalis, Streptococcus sobri-nus, Candida albicans, Actinomyces naeslundii, andFusobacterium nucleatum to alkaline stress. Findingsshowed that planktonic microorganisms were mostsusceptible; only E. faecalis and C. albicans survivedin saturated solution for 10 min (the latter also for100 min). Dentin adhesion was the major factor inimproving the resistance of E. faecalis and A. naeslun-dii to calcium hydroxide, whereas the resistance todisinfecting agents by S. sobrinus was dependent on amulti-species biofilm. In contrast, the C. albicansresponse to calcium hydroxide was not influenced bygrowth conditions. Tolerance to alkali, and possiblyother agents, is likely to be connected to the expres-sion of phenotypes resistant to these agents withinthe biofilm communities. There is very little dataavailable about the effect of calcium hydroxide onbiofilm bacteria. Chavez et al. (141) reported thatbacteria isolated from infected root canals resistedalkaline stress better in biofilms than in planktoniccultures.Endodontic infections are polymicrobial and no

    medicament is effective against all of the bacteriafound in infected root canals. The combination oftwo medicaments may produce additive or synergisticeffects. Evidence suggests that the association ofcalcium hydroxide with CMCP has a broader antibac-terial spectrum, a higher radius of antibacterial action,and kills bacteria faster than mixtures of calciumhydroxide with inert vehicles (142). Therefore,CMCP cannot be considered a vehicle for calciumhydroxide, it is an additional medicament. While invivo studies have indicated calcium hydroxide to be themost effective all-purpose intracanal medicament, IPIand CHX may be able to kill calcium hydroxide re-sistant bacteria. Supplementing the antibacterial activ-ity of Ca(OH)2 with IPI or CHX preparations wasstudied in bovine dentin blocks (143). While Ca(OH)2was unable to kill E. faecalis in the dentin, Ca(OH)2combined with IPI or CHX effectively disinfected thedentin. The addition of CHX or IPI did not affect thealkalinity of the calcium hydroxide suspensions. It maybe assumed that combinations also have the potentialto be used as long-term medication.

    Current therapeutic options for endodontic biofilms

    91

  • Can sealers and cements killbiofilm bacteria?

    An ideal endodontic sealer should be biocompatibleand dimensionally stable; it should seal well and have astrong, long-lasting antimicrobial effect (144146).The antibacterial activity of sealers may help to elimi-nate residual microorganisms that have survivedchemomechanical preparation and thereby improvethe success rate of endodontic treatment. It is expectedthat the antibacterial activity of the root canal sealer, inits unset stage (147), kills the organisms or theybecome deprived of nutritional supply and space forregrowth if pathways to and from the periapical tissueare effectively blocked.It should be recognized that sealers with high anti-

    microbial activity, especially formaldehyde-releasingZnOE (zinc oxideeugenol) sealers such as N2, arealso toxic to cells and tissues. Furthermore, sealers thatrelease antimicrobial substances may also disintegrateto some extent during this stage. Most sealers are onlyantimicrobially active during the setting period (147).For a short time of a few hours or a few days, residualbacteria may be killed. However, this may be enoughto control the residual infection.One of the challenges in endodontic research has

    been the lack of standardized in vitro and in vivoprotocols for testing the antimicrobial effect of sealers.The agar diffusion test (ADT) used to be the mostcommonly applied method to assess the antimicrobialactivity of endodontic sealers (148150). However,the limitations of this method are now well recog-nized. The results obtained are not likely to reflect thetrue antimicrobial potential of the various sealers ordisinfecting agents; therefore, ADT is no longer rec-ommended to be used for this purpose in endodonticresearch (151). A direct contact test (DCT), whichcircumvents many of the problems of ADT, was firstintroduced by Weiss et al. (152) for the evaluation ofthe antimicrobial effect of endodontic sealers and root-end filling materials. The test is a quantitative andreproducible assay that allows for the testing ofinsoluble materials and can be used in standardizedsettings. One study (147) used a modified DCT assayto evaluate the antibacterial activity of seven differentendodontic sealers against E. faecalis 20 minutes aftermixing (fresh samples) and 1, 3, and 7 days aftermixing (set samples). The findings showed that freshiRoot SP, AH Plus, Sealapex, and EndoRez killed

    E. faecalis effectively. iRoot SP, Sealapex, andEndoRez continued to be effective for 3 days aftermixing. Sealapex and EndoRez were the only oneswith continuing antimicrobial activity even at 7 daysafter mixing. However, the direct contact test is notdirectly applicable to studying the effect of the sealeron biofilm bacteria and thus further development ofthe experimental model is warranted.Antibacterial nanoparticulates are found to have

    higher antibacterial activity than antibacterial powders.This is due to the greater surface area and chargedensity of nanoparticulates, which enable them toachieve a higher degree of interaction with the nega-tively charged surface of bacterial cells. Chitosan (CS)is a non-toxic biopolymer derived from the deacetyla-tion of chitin. It is a bioadhesive that readily bindsto negatively charged surfaces and has excellentantimicrobial and antifungal activities. Recently,Kishen et al. (153) examined the ability of differentnanoparticulate-treated dentin to prevent the adher-ence of E. faecalis. Results showed that the incorpora-tion of nanoparticulates did not alter the flowcharacteristics of the ZnOE sealer but improved thedirect antibacterial property and the ability to leachout antibacterial components.

    Eradication of root surface andother extraradicular biofilmsBacteria (or yeasts) that have succeeded in establishinga colony/biofilm on the root surface or in the periapi-cal tissue are beyond the direct reach of conservative(non-surgical) treatment methods. In addition to tra-ditional microbial micro- and macro-colonies, a specialvariant of root surface biofilm is calcified root surfacebiofilm. In vitro, the precipitation of minerals inE. faecalis biofilm on dentin has been described (154),and there are some reports where calculus-like depos-its on the apical external root apex were responsible forroot canal treatment failure (155).In a recent study, Ricucci & Siqueira (16) evaluated

    the prevalence of bacterial biofilms in untreated andtreated root canals of teeth with apical periodontitis.Extraradicular bacterial biofilms were observed in sixout of 100 specimens (6%), four from teeth withuntreated canals and two from teeth with treatedcanals. All of the cases showing an extraradicularbiofilm exhibited clinical symptoms, and three of themwere associated with sinus tracts. These findings indi-

    Haapasalo & Shen

    92

  • cate that extraradicular infections in the form of bio-films or planktonic bacteria are not common.Periapical biofilm colonies, similar to biofilms in

    general, are assumed to be resistant to systemic anti-biotic treatment (156,157). Irrespective of thepathway of infection, when actinomycosis-like coloniesin the tissue or root surface biofilm have developed,surgical treatment including apicoectomy and removalof the infected hard and soft tissues has been shown tobe effective with an excellent long-term prognosis(158160).

    ConclusionsThe complex anatomy of teeth and root canals createsan environment that is a challenge to instrument andclean. In addition, the complex chemical environmentof the root canal prevents antimicrobial irrigating solu-tions and medicaments from exerting their full poten-tial against the microorganisms found in endodonticinfections. While our knowledge of persistent infec-tions, disinfecting agents, and the chemical milieu ofthe necrotic root canal has greatly increased, there is nodoubt that more innovative basic and clinical researchis needed to improve and optimize the use of existingmethods and materials, and to find new techniques andmaterials (or combinations of materials) in order toachieve the goal of predictable, complete disinfectionof the root canal system in apical periodontitis.

    References

    1. Kakehashi S, Stanley HR, Fitzgerald RJ. The effects ofsurgical exposures of dental pulps in germfree andconventional laboratory rats. J South Calif Dent Assoc1966: 34: 449451.

    2. Sundqvist G. Bacteriological studies of necrotic dentalpulps. Umea University Odontological Dissertation.Umea, Sweden: University of Umea, 1976: 7.

    3. Haapasalo M, Udns T, Endal U. Persistent, recur-rent, and acquired infection of the root canal systempost-treatment. Endod Topics 2003: 6: 2956.

    4. Haapasalo M, Endal U, Zandi H, Coil JM. Eradicationof endodontic infection by instrumentation and irriga-tion solutions. Endod Topics 2005: 10: 77102.

    5. Goldman LB, Goldman M, Kronman JH, Lin PS. Theefficacy of several irrigating solutions for endodontics:a scanning electron microscopic study. Oral Surg OralMed Oral Pathol 1981: 52: 197204.

    6. Siqueira JF Jr, Araujo MC, Garcia PF, Fraga RC,Dantas CJ. Histological evaluation of the effectivenessof five instrumentation techniques for cleaning the

    apical third of root canals. J Endod 1997: 23: 499502.

    7. Wu MK, Wesselink PR. A primary observation on thepreparation and obturation of oval canals. Int Endod J2001: 34: 137141.

    8. Tan BT, Messer HH. The quality of apical canal prepa-ration using hand and rotary instruments with specificcriteria for enlargement based on initial apical file size.J Endod 2002: 28: 658664.

    9. Wu MK, van der Sluis LW, Wesselink PR. The capa-bility of two hand instrumentation techniques toremove the inner layer of dentine in oval canals. IntEndod J 2003: 36: 218224.

    10. Peters OA. Current challenges and concepts in thepreparation of root canal systems: a review. J Endod2004: 30: 559567.

    11. Skidmore AE, Bjorndal AM. Root canal morphologyof the human mandibular first molar. Oral Surg OralMed Oral Pathol 1971: 32: 778784.

    12. Vertucci FJ. Root canal anatomy of the human perma-nent teeth. Oral Surg Oral Med Oral Pathol 1984: 58:589599.

    13. Ricucci D, Siqueira JF Jr. Fate of the tissue in lateralcanals and apical ramifications in response to patho-logic conditions and treatment procedures. J Endod2010: 36: 115.

    14. Wu MK, van der Sluis LW, Wesselink PR. A prelimi-nary study of the percentage of gutta-percha-filled areain the apical canal filled with vertically compactedwarm gutta-percha. Int Endod J 2002: 35: 527535.

    15. Gulabivala K, Patel B, Evans G, Ng YL. Effects ofmechanical and chemical procedures on root canalsurfaces. Endod Topics 2005: 10: 103122.

    16. Ricucci D, Siqueira JF Jr. Biofilms and apical periodon-titis: study of prevalence and association with clinicaland histopathologic findings. J Endod 2010: 36:12771288.

    17. Mah TF, OToole GA. Mechanisms of biofilm re-sistance to antimicrobial agents. Trends Microbiol2001: 9: 3439.

    18. Svensater G, Bergenholtz G. Biofilms in endodonticsinfections. Endod Topics 2004: 9: 2736.

    19. Stewart PS. Diffusion in biofilms. J Bacteriol 2003:185: 14851491.

    20. Portenier I, Haapasalo H, Rye A, Waltimo T, rstavikD, Haapasalo M. Inactivation of root canal medica-ments by dentine, hydroxylapatite and bovine serumalbumin. Int Endod J 2001: 34: 184188.

    21. Portenier I, Haapasalo H, rstavik D, Yamauchi M,Haapasalo M. Inactivation of the antibacterial activityof iodine potassium iodide and chlorhexidine diglu-conate against Enterococcus faecalis by dentin, dentinmatrix, type-I collagen, and heat killed microbialwhole cells. J Endod 2002: 28: 634637.

    22. Haapasalo M, Qian W, Portenier I, Waltimo T. Effectsof dentin on the antimicrobial properties of endodon-tic medicaments. J Endod 2007: 33: 917925.

    23. Oliveira JC, Alves FR, Uzeda M, Ras IN, Siqueira JFJr. Influence of serum and necrotic soft tissue on the

    Current therapeutic options for endodontic biofilms

    93

  • antimicrobial effects of intracanal medicaments. BrazDent J 2010: 21: 295300.

    24. Dammaschke T, Witt M, Ott K, Schfer E. Scanningelectron microscopic investigation of incidence, loca-tion, and size of accessory foramina in primary andpermanent molars. Quintessence Int 2004: 35: 699705.

    25. Shen Y, Gao Y, Qian W, Ruse ND, Zhou X, Wu H,Haapasalo M. Three-dimensional numeric simulationof root canal irrigant flow with different irrigationneedles. J Endod 2010: 36: 884889.

    26. Shen Y, Stojicic S, Haapasalo M. Antimicrobial efficacyof chlorhexidine against bacteria in biofilms at differentstages of development. J Endod 2011: 37: 657661.

    27. Dunavant TR, Regan JD, Glickman GN, Solomon ES,Honeyman AL. Comparative evaluation of endodonticirrigants against Enterococcus faecalis biofilms. J Endod2006: 32: 527531.

    28. Williamson AE, Cardon JW, Drake DR. Antimicrobialsusceptibility of monoculture biofilms of a clinicalisolate of Enterococcus faecalis. J Endod 2009: 35:9597.

    29. Shen Y, Qian W, Chung C, Olsen I, Haapasalo M.Evaluation of the effect of two chlorhexidine prepara-tions on biofilm bacteria in vitro: a three-dimensionalquantitative analysis. J Endod 2009: 35: 981985.

    30. Ruddle CJ. Cleaning and shaping the root canalsystem. In: Cohen S, Burns RC, eds. Pathways of thePulp, 8th edn. St. Louis: Mosby Inc., 2002: 231291.

    31. Ando N, Hoshino E. Predominant obligate anaerobesinvading the deep layers of root canal dentin. IntEndod J 1990: 23: 2027.

    32. Peters LB, Wesselink PR, Buijs JF, Van Winkelhoff AJ.Viable bacteria in root dentinal tubules of teeth withapical periodontitis. J Endod 2001: 27: 7681.

    33. Nair PN, Henry S, Cano V, Vera J. Microbial status ofapical root canal system of human mandibular firstmolars with primary apical periodontitis after one-visit endodontic treatment. Oral Surg Oral Med OralPathol Oral Radiol Endod 2005: 99: 231252.

    34. Ricucci D, Siqueira JF Jr. Apical actinomycosis as acontinuum of intraradicular and extraradicular infec-tion: case report and critical review on its involvementwith treatment failure. J Endod 2008: 34: 11241129.

    35. Ferreira FB, Ferreira AL, Gomes BP, Souza-Fiho FJ.Resolution of persistent periapical infection by endo-dontic surgery. Int Endod J 2004: 37: 6169.

    36. Shen Y, Stojicic S, Haapasalo M. Bacterial viability instarved and revitalized biofilms: comparison of viabilitystaining and direct culture. J Endod 2010: 36: 18201823.

    37. Chvez de Paz LE, Hamilton IR, Svenster G. Oralbacteria in biofilms exhibit slow reactivation fromnutrient deprivation. Microbiology 2008: 154: 19271938.

    38. Liu H, Wei X, Ling J, Wang W, Huang X. Biofilmformation capability of Enterococcus faecalis cells instarvation phase and its susceptibility to sodiumhypochlorite. J Endod 2010: 36: 630635.

    39. Siqueira JF Jr, Lima KC, Magalhaes FA, Lopes HP, deUzeda M. Mechanical reduction of the bacterial popu-lation in the root canal by three instrumentation tech-niques. J Endod 1999: 25: 332335.

    40. Dalton BC, rstavik D, Phillips C, Pettiette M, TropeM. Bacterial reduction with nickeltitanium rotaryinstrumentation. J Endod 1998: 24: 763767.

    41. Carver K, Nusstein J, Reader A, Beck M. In vivoantibacterial efficacy of ultrasound after hand androtary instrumentation in human mandibular molars.J Endod 2007: 33: 10381043.

    42. Haapasalo M, rstavik D. In vitro infection and dis-infection of dentinal tubules. J Dent Res 1987: 66:13751379.

    43. Love RM, Jenkinson HF. Invasion of dentinal tubulesby oral bacteria. Crit Rev Oral Biol Med 2002: 13:171183.

    44. El Ayouti A, Chu AL, Kimonis I, Klein C, Weiger R,Lost C. Efficacy of rotary instruments with greatertaper in preparing oval root canals. Int Endod J 2008:41: 10881092.

    45. Paqu F, Ganahl D, Peters OA. Effects of root canalpreparation on apical geometry assessed by micro-computed tomography. J Endod 2009: 35: 10561059.

    46. Peters OA, Schnenberger K, Laib A. Effects of fourNiTi preparation techniques on root canal geometryassessed by micro computed tomography. Int Endod J2001: 34: 221230.

    47. McComb D, Smith DC, Beagrie GS. The results of invivo endodontic chemomechanical instrumentationascanning electron microscope study. J Br Endod Soc1976: 9: 1117.

    48. Metzger Z, Teperovich E, Zary R, Cohen R, Hof R.The self-adjusting file (SAF). Part 1: respecting theroot canal anatomya new concept of endodontic filesand its implementation. J Endod 2010: 36: 679690.

    49. Peters OA, Boessler C, Paqu F. Root canal prepara-tion with a novel nickeltitanium instrument evaluatedwith micro-computed tomography: canal surfacepreparation over time. J Endod 2010: 36: 10681072.

    50. Peters OA, Paqu F. Root canal preparation of maxil-lary molars with the self adjusting file: a micro-computed tomography study. J Endod 2011: 37:5357.

    51. Siqueira JF Jr, Alves FR, Almeida BM, de Oliveira JC,Ras IN. Ability of chemomechanical preparationwith either rotary instruments or self-adjusting file todisinfect oval-shaped root canals. J Endod 2010: 36:18601865.

    52. Hsu Y, Kim S. The resected root surface: the issue ofcanal isthmuses. Dent Clin North Am 1997: 41: 529540.

    53. Endal U, Shen Y, Knut A, Gao Y, Haapasalo M. Ahigh-resolution computed tomographic study ofchanges in root canal isthmus area by instrumentationand root filling. J Endod 2011: 37: 223227.

    54. Burleson A, Nusstein J, Reader A, Beck M. Thein vivo evaluation of hand/rotary/ultrasound

    Haapasalo & Shen

    94

  • instrumentation in necrotic, human mandibularmolars. J Endod 2007: 33: 782787.

    55. Carr GB, Schwartz RS, Schaudinn C, Gorur A, Cos-terton JW. Ultrastructural examination of failed molarretreatment with secondary apical periodontitis: anexamination of endodontic biofilms in an endodonticretreatment failure. J Endod 2009: 35: 13031309.

    56. Duggan JM, Sedgley CM. Biofilm formation of oraland endodontic Enterococcus faecalis. J Endod 2007:33: 815818.

    57. Brndle N, Zehnder M, Weiger R, Waltimo T. Impactof growth conditions on susceptibility of five microbialspecies to alkaline stress. J Endod 2008: 34: 579582.

    58. Chvez de Paz LE, Bergenholtz G, Svenster G. Theeffects of antimicrobials on endodontic biofilm bacte-ria. J Endod 2010: 36: 7077.

    59. Chai WL, Hamimah H, Cheng SC, Sallam AA, Abdul-lah M. Susceptibility of Enterococcus faecalis biofilm toantibiotics and calcium hydroxide. J Oral Sci 2007: 49:161166.

    60. Haapasalo M, Shen Y, Qian W, Gao Y. Irrigation inendodontics. Dent Clin North Am 2010: 54: 291312.

    61. Senia ES, Marshal FJ, Rosen S. The solvent action ofsodium hypochlorite on pulp tissue of extracted teeth.Oral Surg Oral Med Oral Pathol 1971: 31: 96103.

    62. Clegg MS, Vertucci FJ, Walker C, Belanger M, BrittoLR. The effect of exposure to irrigant solutions onapical dentin biofilms in vitro. J Endod 2006: 32:434437.

    63. Haapasalo HK, Siren EK, Waltimo TM, rstavik D,Haapasalo MP. Inactivation of local root canal me-dicaments by dentine: an in vitro study. Int Endod J2000: 33: 126131.

    64. Shaker LA, Dancer BN, Russell AD, Furr JR. Emer-gence and development of chlorhexidine resistanceduring sporulation of Bacillus subtilis 168. FEMSMicrobiol Lett 1988: 51: 7376.

    65. Russell AD, Day MJ. Antibacterial activity of chlor-hexidine. J Hosp Infect 1993: 25: 229238.

    66. Russell AD. Activity of biocides against mycobacteria.Soc Appl Bacteriol Symp Ser 1996: 25: 87S101S.

    67. McDonnell G, Russell AD. Antiseptics and disin-fectants: activity, action, and resistance. Clin MicrobiolRev 1999: 12: 147179.

    68. Vahdaty A, Pitt Ford TR, Wilson RF. Efficacy of chlor-hexidine in disinfecting dentinal tubules in vitro.Endod Dent Traumatol 1993: 9: 243248.

    69. Jeansonne MJ, White RR. A comparison of 2.0% chlor-hexidine gluconate and 5.25% sodium hypochlorite asantimicrobial endodontic irrigants. J Endod 1994: 20:276278.

    70. Heling I, Chandler NP. Antimicrobial effect of irrigantcombinations within dentinal tubules. Int Endod J1998: 31: 814.

    71. Buck RA, Eleazer PD, Staat RH, Scheetz JP. Effec-tiveness of three endodontic irrigants at varioustubular depths in human dentin. J Endod 2001: 27:206208.

    72. Torabinejad M, Khademi AA, Babagoli J, Cho Y,Johnson WB, Bozhilov K, Kim J, Shabahang S. A newsolution for the removal of the smear layer. J Endod2003: 29: 170175.

    73. Torabinejad M, Shabahang S, Aprecio RM, KetteringJD. The antimicrobial effect of MTAD: an in vitroinvestigation. J Endod 2003: 29: 400403.

    74. Shabahang S, Torabinejad M. Effect of MTAD onEnterococcus faecalis contaminated root canals ofextracted human teeth. J Endod 2003: 29: 576579.

    75. Ruff ML, McClanahan SB, Babel BS. In vitro antifun-gal efficacy of four irrigants as a final rinse. J Endod2006: 32: 331333.

    76. Pappen FG, Shen Y, Qian W, Leonardo MR, GiardinoL, Haapasalo M. In vitro antibacterial action of Tetra-clean, MTAD and five experimental irrigation solu-tions. Int Endod J 2010: 43: 528535.

    77. Ma J, Wang Z, Shen Y, Haapasalo M. A new nonin-vasive model to study the effectiveness of dentin dis-infection using confocal laser scanning microscopy.J Endod 2011: 37: 13801385.

    78. Giardino L, Ambu E, Becce C, Rimondini L, MorraM. Surface tension comparison of four common rootcanal irrigants and two new irrigants containing anti-biotic. J Endod 2006: 32: 10911093.

    79. Abou-Rass M, Patonai FJ Jr. The effects of decreasingsurface tension on the flow of irrigating solutions innarrow root canals. Oral Surg Oral Med Oral Pathol1982: 53: 524526.

    80. Stojicic S, Shen Y, Qian W, Johnson B, Haapasalo M.Antibacterial and smear layer removal ability of a novelirrigant, QMiX. Int Endod J 2012: 45: 363371.

    81. Siqueira JF, Uzeda MD. Disinfection by calciumhydroxide pastes of dentinal tubules infected withtwo obligate and one facultative anaerobic bacteria.J Endod 1996: 22: 674676.

    82. George S, Kishen A, Song KP. The role of environ-mental changes on monospecies biofilm formation onroot canal wall by Enterococcus faecalis. J Endod 2005:31: 867872.

    83. rstavik D, Haapasalo M. Disinfection by endodonticirrigants and dressings of experimentally infected den-tinal tubules. Endod Dent Traumatol 1990: 6: 142149.

    84. Kishen A, Sum CP, Mathew S, Lim CT. Influence ofirrigation regimens on the adherence of Enterococcusfaecalis to root canal dentin. J Endod 2008: 34: 850854.

    85. Qian W, Shen Y, Haapasalo M. Quantitative analysis ofthe effect of irrigation sequences on dentin erosion.J Endod 2011: 37: 14371441.

    86. van der Sluis LW, Versluis M, Wu MK, Wesselink PR.Passive ultrasonic irrigation of the root canal: a reviewof the literature. Int Endod J 2007: 40: 415426.

    87. Tronstad L, Barnett F, Schwartzben L, Frasca P. Effec-tiveness and safety of a sonic vibratory endodonticinstrument. Endod Dent Traumatol 1985: 1: 6976.

    Current therapeutic options for endodontic biofilms

    95

  • 88. Ahmad M, Pitt Ford TR, Crum LA. Ultrasonic de-bridement of root canals: an insight into the mech-anisms involved. J Endod 1987: 13: 93101.

    89. Pitt WG. Removal of oral biofilm by sonic phenomena.Am J Dent 2005: 18: 345352.

    90. Caron G. Cleaning Efficiency of the Apical Millimetersof Curved Canals Using Three Different Modalities ofIrrigant Activation: an SEM Study [Masters Thesis].Paris: Paris VII University, 2007.

    91. Brito PR, Souza LC, Machado de Oliveira JC, AlvesFR, De-Deus G, Lopes HP, Siqueira JF Jr. Compari-son of the effectiveness of three irrigation techniquesin reducing intracanal Enterococcus faecalis popula-tions: an in vitro study. J Endod 2009: 35: 14221427.

    92. Huffaker SK, Safavi K, Spangberg LS, Kaufman B.Influence of a passive sonic irrigation system on theelimination of bacteria from root canal systems: a clini-cal study. J Endod 2010: 36: 13151318.

    93. Shen Y, Stojicic S, Qian W, Olsen I, Haapasalo M. Thesynergistic antimicrobial effect by mechanical agitationand two chlorhexidine preparations on biofilm bacte-ria. J Endod 2010: 36: 100104.

    94. Richman MJ. Use of ultrasonics in root canal therapyand root resection. J Dent Med 1956: 12: 12.

    95. Martin H. Ultrasonic disinfection of the root canal.Oral Surg 1976: 42: 9299.

    96. Cunningham WT, Martin H. A scanning electronmicroscope evaluation of root canal debridement withthe endosonic ultrasonic synergistic system. Oral SurgOral Med Oral Pathol 1982: 53: 527531.

    97. Cunningham WT, Martin H, Forrest WR. Evaluationof root canal debridement by the endosonic ultrasonicsynergistic system. Oral Surg Oral Med Oral Pathol1982: 53: 401404.

    98. Collinson KL, Zakariasen MA. Microbiological assess-ment of ultrasonics in root canal therapy. J Endod1986: 12: 131135.

    99. Spoleti P, Siragusa M, Spoleti MJ. Bacteriologicalevaluation of passive ultrasonic activation. J Endod2003: 29: 1214.

    100. Weber CD, McClanahan SB, Miller GA, Diener-WestM, Johnson JD. The effect of passive ultrasonic acti-vation of 2% chlorhexidine or 5.25% sodium hypochlo-rite irrigant on residual antimicrobial activity in rootcanals. J Endod 2003: 29: 562564.

    101. Huque J, Kota K, Yamaga M, Iwaku M, Hoshino E.Bacterial eradication from root dentine by ultrasonicirrigation with sodium hypochlorite. Int Endod J1998: 31: 242250.

    102. Walmsley AD, Laird WR, Lumley PJ. Ultrasound indentistry. Part 2periodontology and endodontics.J Dent 1992: 20: 1117.

    103. Leighton TG. The Acoustic Bubble. London: AcademicPress Limited, 1994.

    104. Lea SC, Price GJ, Walmsley AD. A study to determinewhether cavitation occurs around dental ultrasonicscaling instruments. Ultrason Sonochem 2005: 12:233236.

    105. Fong SW, Klaseboer E, Turangan CK, Khoo BC,Hung KC. Numerical analysis of a gas bubble nearbio-materials in an ultrasound field. Ultrasound MedBiol 2006: 32: 925942.

    106. Hill CR, Bamber JC, ter Haar GR. Physical Principlesof Medical Ultrasonics. West Sussex: John Wiley &Sons Ltd., 2004.

    107. Lumley PJ, Walmsley AD, Laird WR. Streaming pat-terns produced around endosonic files. Int Endod J1991: 24: 290297.

    108. Shrestha A, Fong SW, Khoo BC, Kishen A. Delivery ofantibacterial nanoparticles into dentinal tubules usinghigh-intensity focused ultrasound. J Endod 2009: 35:10281033.

    109. ORiordan K, Sharlin DS, Gross J, Chang S, ErrabelliD, Akilov OE, Kosaka S, Nau GJ, Hasan T. Photoin-activation of Mycobacteria in vitro and in a newmurine model of localized Mycobacterium bovis BCG-induced granulomatous infection. Antimicrob AgentsChemother 2006: 50: 18281834.

    110. Wilson M. Lethal photosensitisation of oral bacteriaand its potential application in the photodynamictherapy of oral infections. Photochem Photobiol Sci2004: 3: 412418.

    111. Bergmans L, Moisiadis P, Huybrechts B, Van Meer-beek B, Quirynen M, Lambrechts P. Effect of photo-activated disinfection on endodontic pathogens exvivo. Int Endod J 2008: 41: 227239.

    112. Bertoloni G, Lauro FM, Cortella G, Merchat M. Pho-tosensitizing activity of hematoporphyrin on Staphylo-coccus aureus cells. Biochim Biophys Acta 2000: 1475:169174.

    113. Romanova NA, Brovko LY, Moore L, Pometun E,Savitsky AP, Ugarova NN, Griffiths MW. Assessmentof photodynamic destruction of Escherichia coliO157:H7 and Listeria monocytogenes by using ATPbioluminescence. Appl Environ Microbiol 2003: 69:63936398.

    114. Nitzan Y, Gutterman M, Malik Z, Ehrenberg B.Inactivation of Gram-negative bacteria by photosensi-tized porphyrins. Photochem Photobiol 1992: 55:8996.

    115. Upadya MH, Kishen A. Influence of bacterial growthmodes on the susceptibility to light-activated disinfec-tion. Int Endod J 2010: 43: 978987.

    116. Leive L. The barrier function of the Gram-negativeenvelope. Ann NY Acad Sci 1974: 235: 109129.

    117. Denyer SP, Maillard JY. Cellular impermeability anduptake of biocides and antibiotics in Gram-negativebacteria. J Appl Microbiol 2002: 92: 35S45S.

    118. Wilson M, Burns T, Pratten J, Pearson GJ. Bacteria insupragingival plaque samples can be killed by low-power laser light in the presence of a photosensitizer.J Appl Bacteriol 1995: 78: 569574.

    119. Merchat M, Bertolini G, Giacomini P, Villanueva A,Jori G. Meso-substituted cationic porphyrins as effi-cient photosensitizers of Gram-positive and Gram-negative bacteria. J Photochem Photobiol B 1996: 32:153157.

    Haapasalo & Shen

    96

  • 120. Williams JA, Pearson GJ, Colles MJ. Antibacterialaction of photoactivated disinfection {PAD} used onendodontic bacteria in planktonic suspension and inartificial and human root canals. J Dent 2006: 34:363371.

    121. Soukos NS, Chen PSY, Morris JT, Ruggiero K, Aber-nethy AD, Som S, Foschi F, Doucette S, BammannLL, Fontana CR, Doukas AG, Stashenko PP. Photo-dynamic therapy for endodontic disinfection. J Endod2006: 32: 979984.

    122. Fimple JL, Fontana CR, Foschi F, Ruggiero K, SongX, Pagonis TC, Tanner AC, Kent R, Doukas AG,Stashenko PP, Soukos NS. Photodynamic treatment ofendodontic polymicrobial infection in vitro. J Endod2008: 34: 728734.

    123. George S, Kishen A. Photophysical, photochemical,and photobiological characterization of methyleneblue formulations for light-activated root canal disin-fection. J Biomed Opt 2007: 12: 034029.

    124. Foschi F, Fontana CR, Ruggiero K, Riahi R, Vera A,Doukas AG, Pagonis TC, Kent R, Stashenko PP,Soukos NS. Photodynamic inactivation of Enterococcusfaecalis in dental root canals in vitro. Lasers Surg Med2007: 39: 782787.

    125. George S, Kishen A. Advanced noninvasive light-activated disinfection: assessment of cytotoxicity onfibroblast versus antimicrobial activity against Entero-coccus faecalis. J Endod 2007: 33: 599602.

    126. George S, Kishen A. Augmenting the anti-biofilm effi-cacy of advanced noninvasive light activated disinfec-tion with emulsified oxidizer and oxygen carrier. JEndod 2008: 34: 11191123.

    127. George S, Kishen A. Influence of photosensitizersolvent on the mechanisms of photoactivated killing ofEnterococcus faecalis. Photochem Photobiol 2008: 84:734740.

    128. Lima JP, Sampaio de Melo MA, Borges FM, TeixeiraAH, Steiner-Oliveira C, Nobre Dos Santos M, Rod-rigues LK, Zanin IC. Evaluation of the antimicrobialeffect of photodynamic antimicrobial therapy in an insitu model of dentine caries. Eur J Oral Sci 2009: 117:568574.

    129. Fava LR, Saunders WP. Calcium hydroxide pastes:classification and clinical indications. Int Endod J1999: 32: 257282.

    130. Siqueira JF Jr, Lopes HP. Mechanisms of antimicrobialactivity of calcium hydroxide: a critical review. IntEndod J 1999: 32: 361369.

    131. Bystrm A, Sundqvist G. The antibacterial action ofsodium hypochlorite and EDTA in 60 cases of endo-dontic therapy. Int Endod J 1985: 18: 3540.

    132. Hasselgren G, Olsson B, Cvek M. Effects of calciumhydroxide and sodium hypochlorite on the dissolutionof necrotic porcine muscle tissue. J Endod 1988: 14:125127.

    133. Safavi KE, Spngberg LS, LangelandK. Root canal den-tinal tubule disinfection. J Endod 1990: 16: 207210.

    134. Distel JW, Hatton JF, Gillespie MJ. Biofilm formationin medicated root canals. J Endod 2002: 28: 689693.

    135. Sirn EK, Haapasalo MP, Ranta K, Salmi P, KerosuoEN. Microbiological findings and clinical treatmentprocedures in endodontic cases selected for microbio-logical investigation. Int Endod J 1997: 30: 9195.

    136. Waltimo TM, Sirn EK, Torkko HL, Olsen I,Haapasalo MP. Fungi in therapy-resistant apical peri-odontitis. Int Endod J 1997: 30: 96101.

    137. Shuping GB, rstavik D, Sigurdsson A, Trope M.Reduction of intracanal bacteria using nickeltitaniumrotary instrumentation and various medications. JEndod 2000: 26: 751755.

    138. rstavik D, Kerekes K, Molven O. Effects of extensiveapical reaming and calcium hydroxide dressing on bac-terial infection during treatment of apical periodonti-tis: a pilot study. Int Endod J 1991: 24: 17.

    139. Kvist T, Molander A, Dahln G, Reit C. Microbiologi-cal evaluation of one- and two-visit endodontic treat-ment of teeth with apical periodontitis: a randomized,clinical trial. J Endod 2004: 30: 572576.

    140. Peters LB, Wesselink PR. Periapical healing of endo-dontically treated teeth in one and two visits obturatedin the presence or absence of bacteria in the root canal.Int Endod J 2002: 35: 660667.

    141. Chvez de Paz LE, Bergenholtz G, Dahln G, Sven-ster G. Response to alkaline stress by root canal bac-teria in biofilms. Int Endod J 2007: 40: 344355.

    142. Bystrm A, Claesson R, Sundqvist G. The antibacterialeffect of camphorated paramonochlorophenol, cam-phorated phenol and calcium hydroxide in the treat-ment of infected root canals. Endod Dent Traumatol1985: 1: 170175.

    143. Sirn EK, Haapasalo MP, Waltimo TM, rstavik D. Invitro antibacterial effect of calcium hydroxide com-bined with chlorhexidine or iodine potassium iodideon Enterococcus faecalis. Eur J Oral Sci 2004: 112:326331.

    144. Grossman L. Antimicrobial effect of root canalcements. J Endod 1980: 6: 594597.

    145. rstavik D. Antibacterial properties of endodonticmaterials. Int Endod J 1988: 21: 161169.

    146. Geurtsen W, Leyhausen G. Biological aspects of rootcanal filling materials: histocompatibility, cytotoxicity,and mutagenicity. Clin Oral Investig 1997: 1: 511.

    147. Zhang H, Shen Y, Ruse ND, Haapasalo M. Antibac-terial activity of endodontic sealers by modified directcontact test against Enterococcus faecalis. J Endod2009: 35: 10511055.

    148. Siqueira JF Jr, Favieri A, Gahyva SM, Moraes SR, LimaKC, Lopes HP. Antimicrobial activity and flow rate ofnewer and established root canal sealers. J Endod 2000:26: 274277.

    149. Cobankara FK, Altinoz HC, Ergani O, Kav K, Belli S.In vitro antibacterial activities of root-canal sealers byusing two different methods. J Endod 2004: 30:5760.

    150. Sipert CR, Hussne RP, Nishiyama CK, Torres SA. Invitro antimicrobial activity of Fill Canal, Sealapex,Mineral Trioxide Aggregate, Portland cement andEndoRez. Int Endod J 2005: 38: 539543.

    Current therapeutic options for endodontic biofilms

    97

  • 151. Editorial Board of the Journal of Endodontics.Wanted: a base of evidence. J Endod 2007: 33: 14011402.

    152. Weiss EI, Shalhav M, Fuss Z. Assessment of antibac-terial activity of endodontic sealers by a direct contacttest. Endod Dent Traumatol 1996: 12: 179184.

    153. Kishen A, Shi Z, Shrestha A, Neoh KG. An investiga-tion on the antibacterial and antibiofilm efficacy ofcationic nanoparticulates for root canal disinfection.J Endod 2008: 34: 15151520.

    154. Kishen A, George S, Kumar R. Enterococcus faecalismediated biomineralized biofilm formation on rootcanal dentine in vitro. J Biomed Mater Res A 2006: 77:406415.

    155. Ricucci D, Martorano M, Bate AL, Pascon EA.Calculus-like deposit on the apical external rootsurface of teeth with post-treatment apical periodon-titis: report of two cases. Int Endod J 2005: 38: 262271.

    156. Hall-Stoodley L, Stoodley P. Evolving concepts inbiofilm infections. Cell Microbiol 2009: 11: 10341043.

    157. Hoiby N, Bjarnsholt T, Givskov M, Molin S, Ciofu O.Antibiotic resistance of bacterial biofilms. Int J Anti-microb Agents 2010: 35: 322332.

    158. Wesley RK, Osborn TP, Dylewski JJ. Periapical actin-omycosis: clinical considerations. J Endod 1977: 3:352355.

    159. Sundqvist G, Reuterving C. Isolation of Actinomycesisraelii from periapical lesion. J Endod 1980: 6: 602606.

    160. Rubinstein RA, Kim S. Short-term observation of theresults of endodontic surgery with the use of a surgicaloperation microscope and Super-EBA as root-endfilling material. J Endod 1999: 25: 4348.

    Haapasalo & Shen

    98