+ All Categories
Home > Documents > Therapeutic effects of novel resin bonding systems containing bioactive glasses on mineral-depleted...

Therapeutic effects of novel resin bonding systems containing bioactive glasses on mineral-depleted...

Date post: 25-Aug-2016
Category:
Upload: manuel
View: 212 times
Download: 0 times
Share this document with a friend
12
Therapeutic effects of novel resin bonding systems containing bioactive glasses on mineral-depleted areas within the bonded-dentine interface Salvatore Sauro Raquel Osorio Timothy F. Watson Manuel Toledano Received: 27 November 2011 / Accepted: 27 February 2012 / Published online: 1 April 2012 Ó Springer Science+Business Media, LLC 2012 Abstract This study aimed in evaluating the effects of two experimental resin bonding systems containing con- ventional Bioglass 45S5 (BAG) or Zinc-polycarboxylated bioactive glass (BAG-Zn) micro-fillers on the resin-bonded dentine interface after storage in a simulated body fluid solution (SBFS). Three resin bonding systems were for- mulated: Resin-A: (BAG containing); Resin-B; (BAG-Zn containing); Resin-C (no filler). The ability of the experi- mental resins to evoke apatite formation was evaluated using confocal Raman spectroscopy. Acid-etched dentine specimens were bonded, and prepared for AFM/nano- indentation analysis in a fully-hydrated status to evaluate the modulus of elasticity (Ei) and hardness (Hi) across the interface at different SBFS storage periods. Further resin– dentine specimens were tested for microtensile bond strength after 24 h or 3 months of SBFS storage. SEM examination was performed after de-bonding and confocal laser microscopy was used to evaluate the ultramorphology of the interfaces and micropermeability. The resin A and B showed a consistent presence of apatite (967 cm -1 ), reduced micropermeability within the resin–dentine inter- face and a significant increase of the Ei and Hi along the bonded-dentine interface after prolonged SBFS storage. Bond strength values were affected by the resin system (P \ 0.0001) and by storage time (P \ 0.0001) both after 24 h and 3 months of SBFS storage. In conclusion, resin bonding systems containing bioactive fillers may a have therapeutic effect on the nano-mechanical properties and sealing ability of mineral-depleted resin–dentine interface. 1 Introduction The development of innovative bioactive ion-releasing restorative materials with a therapeutic ability to remineralise mineral-depleted sites within the bonded-dentine interface is currently one of the main targets of the dental biomaterial research [15]. In vitro [6, 7] and in vivo [3, 4] studies demonstrated that remineralisation may be achieved by increasing the mineral content and the mechanical properties (i.e. hardness and modules of elasticity) of dental hard tissues. Peters et al. [3] showed by electron probe elemental micro- analysis (EPMA) techniques, an increase mineral content in the caries affected dentine using resin-based materials con- taining calcium phosphate cements. Moreover, specific restorative materials such as glass ionomer cements (GIC) or resin-modified glass ionomer cements (RMGIC) may also promote mineral precipitation in residual caries affected dentine [5, 8, 9]. Ryou et al. [8] have recently demonstrated that the remineralisation of such demineralised zones within the hybrid layers may be a potential means for preserving the durability of the resin–dentine interface [8, 12]. Indeed, poorly resin-infiltrated hybrid layers created in acid-etched dentine may be characterised by collagen degradation within a resin-sparse demineralised zone which frequently remains subsequent to bonding procedures [10, 11]. However, the use of Zinc-containing materials may reduce the collagen deg- radation mediated by matrix metalloproteinases (MMPs) within the poorly infiltrated hybrid layers and carious dentine by protecting sensitive cleavage sites of collagen within the demineralised dentine [13]. S. Sauro (&) Á T. F. Watson Biomaterials, Biomimetics and Biophotonics (B3 group), King’s College London Dental Institute, Guy’s Dental Hospital, London SE1 9RT, England, UK e-mail: [email protected] S. Sauro Á R. Osorio Á M. Toledano Dental Materials, School of Dentistry, University of Granada, Colegio Ma ´ximo, Campus de Cartuja, Granada, Spain 123 J Mater Sci: Mater Med (2012) 23:1521–1532 DOI 10.1007/s10856-012-4606-6
Transcript

Therapeutic effects of novel resin bonding systemscontaining bioactive glasses on mineral-depleted areaswithin the bonded-dentine interface

Salvatore Sauro • Raquel Osorio • Timothy F. Watson •

Manuel Toledano

Received: 27 November 2011 / Accepted: 27 February 2012 / Published online: 1 April 2012

� Springer Science+Business Media, LLC 2012

Abstract This study aimed in evaluating the effects of

two experimental resin bonding systems containing con-

ventional Bioglass 45S5 (BAG) or Zinc-polycarboxylated

bioactive glass (BAG-Zn) micro-fillers on the resin-bonded

dentine interface after storage in a simulated body fluid

solution (SBFS). Three resin bonding systems were for-

mulated: Resin-A: (BAG containing); Resin-B; (BAG-Zn

containing); Resin-C (no filler). The ability of the experi-

mental resins to evoke apatite formation was evaluated

using confocal Raman spectroscopy. Acid-etched dentine

specimens were bonded, and prepared for AFM/nano-

indentation analysis in a fully-hydrated status to evaluate

the modulus of elasticity (Ei) and hardness (Hi) across the

interface at different SBFS storage periods. Further resin–

dentine specimens were tested for microtensile bond

strength after 24 h or 3 months of SBFS storage. SEM

examination was performed after de-bonding and confocal

laser microscopy was used to evaluate the ultramorphology

of the interfaces and micropermeability. The resin A and B

showed a consistent presence of apatite (967 cm-1),

reduced micropermeability within the resin–dentine inter-

face and a significant increase of the Ei and Hi along the

bonded-dentine interface after prolonged SBFS storage.

Bond strength values were affected by the resin system

(P \ 0.0001) and by storage time (P \ 0.0001) both after

24 h and 3 months of SBFS storage. In conclusion, resin

bonding systems containing bioactive fillers may a have

therapeutic effect on the nano-mechanical properties and

sealing ability of mineral-depleted resin–dentine interface.

1 Introduction

The development of innovative bioactive ion-releasing

restorative materials with a therapeutic ability to remineralise

mineral-depleted sites within the bonded-dentine interface is

currently one of the main targets of the dental biomaterial

research [1–5]. In vitro [6, 7] and in vivo [3, 4] studies

demonstrated that remineralisation may be achieved by

increasing the mineral content and the mechanical properties

(i.e. hardness and modules of elasticity) of dental hard tissues.

Peters et al. [3] showed by electron probe elemental micro-

analysis (EPMA) techniques, an increase mineral content in

the caries affected dentine using resin-based materials con-

taining calcium phosphate cements. Moreover, specific

restorative materials such as glass ionomer cements (GIC) or

resin-modified glass ionomer cements (RMGIC) may also

promote mineral precipitation in residual caries affected

dentine [5, 8, 9]. Ryou et al. [8] have recently demonstrated

that the remineralisation of such demineralised zones within

the hybrid layers may be a potential means for preserving the

durability of the resin–dentine interface [8, 12]. Indeed,

poorly resin-infiltrated hybrid layers created in acid-etched

dentine may be characterised by collagen degradation within

a resin-sparse demineralised zone which frequently remains

subsequent to bonding procedures [10, 11]. However, the use

of Zinc-containing materials may reduce the collagen deg-

radation mediated by matrix metalloproteinases (MMPs)

within the poorly infiltrated hybrid layers and carious dentine

by protecting sensitive cleavage sites of collagen within the

demineralised dentine [13].

S. Sauro (&) � T. F. Watson

Biomaterials, Biomimetics and Biophotonics (B3 group), King’s

College London Dental Institute, Guy’s Dental Hospital, London

SE1 9RT, England, UK

e-mail: [email protected]

S. Sauro � R. Osorio � M. Toledano

Dental Materials, School of Dentistry, University of Granada,

Colegio Maximo, Campus de Cartuja, Granada, Spain

123

J Mater Sci: Mater Med (2012) 23:1521–1532

DOI 10.1007/s10856-012-4606-6

Bioglass 45S5, a high biocompatible calcium/sodium

phosphate-phyllosilicate originally developed as a bone

conductive material but chemically similar to natural tooth

mineral, is one of the foremost bioactive materials with an

excellent ability to induce calcium phosphate (Ca/P) pre-

cipitation and subsequent crystallisation into hydroxyapa-

tite [HA: Ca10(PO4)6(OH)2] when immersed in protein-

free simulated body fluid solution (SBFS) or saliva

[14]. Whereas Bioglass 45S5 has been successfully used

for dentine remineralisation when directly applied on

demineralised dentinal tissue (16), the development of resin-

based restorative materials containing bioactive micro-fill-

ers with remineralising effects on the mineral-depleted areas

within the bonded-dentine interface remains an important

target to accomplish.

The aim of this study was to evaluate the ability of two

experimental resin bonding systems containing Bioglass

45S5 (BAG) or Zinc-polycarboxylated bioactive glass

(BAG-Zn) to induce therapeutic effects on the bonded-

dentine interface after SBFS ageing. The null hypothesis to

be tested is that the inclusion of bioactive micro-fillers

within the composition of the resin bonding systems

encourages no remineralisation on the mineral-depleted

areas within the bonded-dentine interface.

2 Materials and methods

2.1 Bioactive fillers and experimental bonding systems

preparation

A BAG micro-filler was produced by melting 46.1 mol%

SiO2, 26.9 mol% CaO, 24.4 mol% Na2O and 2.5 mol%

P2O5 in a platinum crucible at 10�C/min up to 1,100�C and

maintained for 1 h. Subsequently the temperature was raised

to 1450�C (10�C/min) and maintained for further 30 min.

The glass melt was rapidly quenched in water (~20�C),

dehydrated in absolute ethanol for 3 h, dried overnight in a

furnace at 110�C, milled and finally sieved (\20 lm). The

BAG-Zn micro-filler was produced by mixing 80 wt% BAG

with 20 wt% Zinc oxide (Aldrich Chemical Co, Gillingham,

UK) and mixed (ratio 1:2) with 10% polycarboxylic acid

solution (PAA: MW 1800; Aldrich Chemical Co.) for 5 min

(40 rpm). The mixture was dried in a furnace at 80�C for

12 h. The final product was finally milled and sieved as

previously described. A resin co-monomer blend was for-

mulated using three hydrophobic monomers (UDMA,

BisGMA, TEGDMA: Esstech Essington, PA, USA) a

hydrophilic monomer (HEMA) and absolute ethanol

(Aldrich Chemical Co.). The resin solution was made light-

curable by adding 0.5 wt% of camphoroquinone and

1.0 wt% of ethyl 4-dimethylaminobenzoate (Aldrich

Chemical Co.). 5-min sonication (Model QS3, Ultrawave

Ltd, Cardiff, UK) and 2-day shaking (Orbital Shakers PSU-

20i, Cole Fisher Scientific Ltd, Loughborough, UK) were

required to achieve a well-mixed resin blend. Three exper-

imental resin bonding systems were subsequently created:

(1) Resin A: 60 wt% resin blend/40 wt% BAG; (2) Resin B:

60 wt% resin blend/40 wt% BAG-Zn; (3) Resin-C: 100%

resin solution (no filler). The detailed composition of each

resin bonding system is shown in Table 1.

2.2 Confocal Raman spectroscopy evaluation

Three disks for each resin were created using a TEFLON

cylinder mold (8 mm in dm and 1 mm thick) and a light-

curing system (600 mWcm2, Optilux VLC, Demetron,

Research Co., CT, USA) for 1 min. A computer-controlled

confocal laser Raman apparatus equipped with a Leica

DM/LM optical microscope with a 209/0.40 NA objective

and CCD detector attached to a modular research spec-

trograph (Renishaw InVia; Renishaw plc, Gloucheshire,

UK) was used to evaluate the apatite precipitation after

24 h, 1 month (1 m) and 3 months (3 m) of SBFS aging. A

near-infrared diode laser spot size (\1 lm) operating at

785 nm was used to induce the Raman scattering effect

(200–1,100 cm-1). The entire surface of the specimens

was examined with steps of 20.0 lm on the X and Y axes

using a computer-motorised stage and analysed for the

peak of apatite at 967 cm-1 using the software Wire 3.2

(Renishaw), [15]. A SBF solution containing NaCl,

NaHCO3, KCl, K2HPO4, MgCl2, CaCl2, and (CH2OH)3

CNH2/HCl (6 N) to buffer the pH (7.4) was replaced every

72 h.

2.3 Dentine specimen preparation and bonding

procedures

Human caries-free molars extracted for surgical reasons

under a protocol approved by an institutional review board

were stored in deionised water (pH 7.1) at 4�C for no

longer than one 1 month. Dentine crown specimens were

prepared by cutting the roots 1 mm beneath the cemento–

enamel junction (CEJ) using a hard tissue microtome

(Isomet 11/1180, Buehler, Coventry, UK) equipped with a

diamond embedded blade (XL 12205, Benetec Limited,

London, UK). The occlusal enamel was then removed with

a second parallel cut in order to expose a middle coronal

dentine. A standard and more clinically relevant smear

layer was created using 180-grit SiC paper for 30 s under

constant water irrigation. The specimens were divided into

three groups based on the tested bonding systems (A, B, C)

and conditioned using a 35% orthophosphoric acid solution

(H3PO4; Aldrich Chemical, Co.) for 15 s followed by a

copious water rinse (5 s). The acid-etched specimens were

air-dried for 2 s, bonded with the experimental adhesives

1522 J Mater Sci: Mater Med (2012) 23:1521–1532

123

within a period of 20 s and immediately light-cured for

30 s (600 mW cm2, Optilux VLC, Demetron, Research

Co., CT, USA). A flowable resin composite (X-FlowTM,

Dentsply, Caulk, UK) was placed incrementally in two

1 mm layers and light-cured for 40 s. The detailed appli-

cation mode is shown in Table 1.

2.4 AFM imaging and nano-indentation

Three bonded-dentine specimens for each group were lon-

gitudinally sectioned in slabs (Thickness: 1.5 mm) and

polished through SiC abrasive papers from 800 up to

4000-grit. A final polishing procedure was performed using

diamond pastes (Buheler-MetaDi, Buheler Ltd. Lake Bluff,

IL, USA) through 1 lm down to 0.25 lm. The specimens

were treated in ultrasonic bath (Model QS3, Ultrawave Ltd,

Cardiff, UK) containing deionised water (pH 7.4) for 5 min

at each polishing step. An atomic force microscope (AFM -

Nanoscope V, Digital Instruments, Veeco Metrology group,

Santa Barbara, CA, USA) equipped with a Triboscope

indentor system (Hysitron Inc., Minneapolis, MN) and a

water immersion Berkovich-tip indenter (radius *20 nm)

was employed for the imaging and indentation processes

in a fully-hydrated status. Six indentations with a load of

4000 nN and a time function of 10s were performed in a

straight line starting from the adhesive layer down to the

intertubular dentine at different SBFS storage periods (24 h,

1 m, 3 m) in order to evaluate the modulus of Ei and Hi.

Three indentation lines were executed in five different

mesio-distal positions along the interface. The distance

between each indentation was kept constant by adjusting the

distance intervals in 5 (±1) lm steps and the load function

[16, 17]. ANOVA was performed including Ei or Hi as a

dependent variable. Resin adhesive type (A/B/C), period of

SBFS storage (24 h/1 m/3 m), and position along the

interface (1st–6th) were considered as independent vari-

ables. Analysis of interaction was also executed. Student–

Newman–Keuls test was used for multiple compari-

sons. Statistical analysis was set at a significance level of

a = 0.05.

2.5 Confocal laser scanning microscopy evaluation

(CLSM)

Three acid-etched dentine specimens for each group were

bonded, as previously described, with the experimental

resins doped with 0.2 wt% Sulfo-Rhodamine 101 (SR-101:

Sigma Chemicals, St. Louis, MO, USA). The pulp chamber

of the specimens was exposed and half of the specimens

were immediately filled with 0.1 wt% water solution of

Lucifer Yellow CH (LY: Sigma Chemicals) for 3 h

[17, 18] while, the remnant half part was immersed in

SBFS for 3 m and subsequently filled with LY. The

specimens were washed with deionised water in an ultra-

sonic bath for 2 min, sectioned into slabs (1.5 mm) and

finally polished using 4000-grit SiC. A confocal laser

Table 1 Composition of the experimental resin bonding system used in this study

Resin Chemical composition Application mode and bonding procedures

Resin A 35% UDMA

3.6% Bis-GMA

3.4% TEGDMA

15% HEMA

10% Absolute ethanol

33% Bioglass 45S5

• Etching with 35%-H3PO4 for 15s

• Abundant rinse with deionised water

• Air-dried for 2 s,

• Application of a first layer of each experimental adhesives for 10 s

• Air-dried for 5s at maximum stram power

• Application of a second layer of each experimental adhesives for 10 s

• Gently air-dried for 2 s,

• Light-cure for 30 s

• Apply resin composite and light-cure

Resin B 35% UDMA

3.6% Bis-GMA

3.4% TEGDMA

15% HEMA

10% Absolute ethanol

33% Zn-polycarboxilated bioactive glass

Resin C 55% UDMA

4.5% Bis-GMA

10.5% TEGDMA

18% HEMA

12% Absolute ethanol

0.5 wt% of camphoroquinone (CQ) and 1.0 wt% of ethyl 4-dimethylaminobenzoate (EDMB) were also included within the composition of the resins

Abbreviations: BisGMA 2,2-bis[4-(2-hydroxy-3-methacryloylpropoxy)]-phenyl propane, TEGDMA triethyleneglycol dimethacrylate, HEMA2-hydroxyethylmethacrylate, UDMA urethane dimethacrylates, Absolute ethanol: 100% ethylic alcohol

J Mater Sci: Mater Med (2012) 23:1521–1532 1523

123

scanning microscope (CLSM: Leica SP2 CLSM, Heidel-

berg, Germany) equipped with a 63x/1.4 NA oil immersion

lens, a 488 nm argon/helium and a 633 nm krypton ion

laser illumination was used to analysed the ultra-mor-

phology and the micropermeability along the resin–dentine

interfaces. Reflection and fluorescence optical images were

captured 5 lm below the outer surface up to 25 lm depth

(1 lm z-step) and converted in topographic single projec-

tions using the Leica SP2 CLSM image-processing soft-

ware (Leica, Heidelberg, Germany). The configuration of

the system was standardised and used at the same level for

the entire investigation.

2.6 Micro-tensile bond strength (lTBS)

Further five bonded-dentine specimens for each group were

sectioned in both X and Y directions to obtain match-sticks

of 0.9 mm2. They were tested after 24 h or 3 m of SBFS

storage using a customised microtensile jigs on a linear

actuator (SMAC Europe Ltd., Horsham, West Sussex, UK).

Mean bond strength values were analysed by two-way

ANOVA and Student–Newman–Keuls multiple compari-

sons. Statistical analysis was performed at a significance

level of a = 0.05. Modes of failure were classified as per-

centage of adhesive (A) or mixed (M) or cohesive (C) using

a 30x USB digital-microscope (CY-800B). Five represen-

tative fractured specimens for each group were sputter-

coated with gold (SCD 004 Sputter Coater; Bal-Tec, Vaduz,

Liechtenstein) and examined with SEM (S-3500; Hitachi,

Wokingham, UK) with an accelerating voltage of 15 kV

and a working distance of 15 mm.

3 Results

3.1 Confocal Raman spectroscopy evaluation

No presence of apatite was detected after 24 h of SBFS

storage but Resin A showed a consistent presence of a

prominent peak at 967 cm-1 both after 1 m and 3 m of

SBFS storage. Conversely, Resin B presented an irregular

detection of apatite after 1 m; a prominent and reliable

peak at 967 cm-1 was detected after 3 m of SBFS storage.

No apatite formation was identified on the specimens cre-

ated using the Resin C after 3 m of SBFS storage (Fig. 1).

3.2 Nano-mechanical properties and AFM/CLSM

microscopy

The nano-mechanical properties (Ei and Hi) were influenced

by resin adhesive type, position along the interface (1st–6th)

and storage time (P \ 0.01), (Table 2 and Fig. 2). Interac-

tions between factors were also significant (P \ 0.001). In

detail, the first indentation performed on the adhesive layer of

the Resin A showed a significant reduction in Ei and Hi after

prolonged SBFS storage (1 m and 3 m) (Fig. 2A, B). Con-

versely, both the second indentation values corresponding to

the hybrid layer (HL) and the third value performed in the

bottom of the HL showed a statistical increase of the Ei and

Hi (P \ 0.01). No significant changes were observed along

the intertubular dentine in the 4th, 5th and 6th position sub-

sequent to prolonged SBFS storage. The AFM imaging

showed a scarcely pronounced indentation mark on the HL

after 24 h of SBFS storage (Fig. 3A). On the contrary, a well

shaped indentation mark was observed both after 1 m

(Fig. 3B) and 3 m (Fig. 3C) of SBFS storage. The confocal

microscopy showed after 24 h an intense micropermeability

within the HL after 24 h (Fig. 3D, E) while, a strong

reflection signal was detected subsequent to 3 m of SBFS

storage (Fig. 3F). The experimental Resin B showed higher

Ei and Hi (P \ 0.01) (Fig. 2C, D) and a pronounced inden-

tation mark both on the HL and bottom of HL after prolonged

SBFS storage (Fig. 4A). The bonded-dentine interface cre-

ated with Resin B (Fig 4B) was affected by important

micropermeability at 24 h (Fig. 4C) which appeared limited

and localised at the bottom of a very reflective HL after 3 m

of SBFS storage (Fig 4D). Resin C showed a statistical

reduction in Ei and Hi (P \ 0.01) of the adhesive layer, HL

and bottom of HL after prolonged SBFS storage (Fig. 2E, F);

a severe micropermeability (Fig. 5A) and an evident gap

were observed within the interface (Fig. 5B).

Fig. 1 Confocal Raman spectroscopy of the resin bonding systems

tested in this study. The resins A, B and C show no peak at 967 cm-1

(apatite) after 24 h of SBFS storage. The outer surfaces of resin A and

B were characterised by a constant presence of apatite after prolonged

SBFS storage (3 m)

1524 J Mater Sci: Mater Med (2012) 23:1521–1532

123

3.3 Micro-tensile bond strength (lTBS) and failure

mode analysis

Bond strength values were affected by the resin type

(F = 8.50; P \ 0.0001) and by storage time (F = 182.58;

P \ 0.0001). Interactions were also significant (P \0.001), (Table 3). Reliability of the model was 77%. In

detail, resin A showed lTBS values up to 44.4 ± 7.4 MPa

and generally failed in a mixed mode after 24 h. The

prolonged SBFS storage (3 m) induced in these dentine-

bonded specimens a significant lTBS drop (P \ 0.05) and

a failure mode prevalently mixed and adhesive. However,

when the 3 m SBFS-aged match-sticks failed in an adhe-

sive mode (Fig. 6A1) were analysed using SEM it was

possible to observe a surface still covered by residual resin

(Fig. 6B1) and mineral crystals embedded in a resin/col-

lagen scaffold-like structure (Fig. 6B2). In contrast, the

match-sticks tested after 24 h showed an exposed dentine

surface with a few crystals only inside the dentinal tubules

(Fig. 6B3). The match-sticks created with the resin B

achieved lTBS values up to 36.6 ± 11.3 MPa after 24 h of

SBFS storage and failed generally in the mixed mode.

Resin B showed a significant lTBS reduction after pro-

longed SBFS storage (3 m) with a failure mode mixed and

adhesive remnant resin and mineral crystals was detected

by SEM on the fractured surfaces (Fig. 6C1, C2). Mineral

crystals were observed only inside the dentinal tubules in

specimens aged for 24 h as shown in Fig. 6B4. The resin C

attained lTBS values up to 37.4 ± 14.7 MPa and failed

prevalently in mixed mode after 24 h. These values sig-

nificantly dropped subsequent to prolonged SBFS storage

(3 m) and the fractured surface was characterised by the

presence of numerous exposed dentinal tubules (Fig. 6D1)

and collagen fibrils (Fig. 6D2–3).

4 Discussion

The nano-mechanical properties and the ultramorphology/

micropermeability of the bonded-dentine interface were

influenced considerably by the experimental resins sub-

sequent to SBFS storage. Therefore, the null hypothesis

that the inclusion of bioactive micro-fillers within the

composition of the resin-based restorative materials has no

Table 2 Mean and standard deviation of AFM nano-mechanical properties (Ei/Hi) along the resin–dentine interface

Ei–24 h Ei–1 month Ei 3–months Hi–24 h Hi–1 month Hi–3 months

Resin A

1st 3.22 ± 1.10 A1 2.30 ± 0.70 A2 1.85 ± 0.40 A2 0.17 ± 0.03 A1 0.11 ± 0.02 A2 0.07 ± 0.03 A2

2nd 1.51 ± 0.48 B1 2.17 ± 0.67 A2 2.35 ± 0.70 B2 0.03 ± 0.01 B1 0.09 ± 0.01 A2 0.09 ± 0.03 A2

3rd 8.40 ± 2.43 C1 10.3 ± 1.58 B2 11.5 ± 2.66 C2 0.37 ± 0.10 C1 0.48 ± 0.11 B2 0.57 ± 0.10 B3

4th 15.9 ± 1.95 D1 16.3 ± 1.38 C1 16.70 ± 0.95 D1 0.73 ± 0.12 D1 0.72 ± 0.06 C1 0.72 ± 0.06 C1

5th 16.9 ± 0.85 DE1 18.0 ± 1.44 D1 17.7 ± 0.82 DE1 0.75 ± 0.06 D1 0.73 ± 0.07 C1 0.73 ± 0.07 C1

6th 17.7 ± 1.30 E1 18.5 ± 0.96 D1 18.8 ± 0.70 E1 0.85 ± 0.08 E1 0.89 ± 0.09 D1 0.89 ± 0.09 D1

Resin B

1st 2.33 ± 0.47 A1 1.71 ± 0.84 A1 1.53 ± 0.53 A2 0.33 ± 0.04 A1 0.21 ± 0.12 A2 0.20 ± 0.14 A2

2nd 1.68 ± 0.50 B1 1.77 ± 0.37 A1 3.17 ± 0.95 B2 0.04 ± 0.01 B1 0.07 ± 0.02 B2 0.09 ± 0.01 B2

3rd 7.43 ± 1.91 C1 9.71 ± 1.79 B2 12.7 ± 1.35 C3 0.47 ± 0.11 C1 0.43 ± 0.11 C1 0.60 ± 0.09 C2

4th 16.0 ± 2.45 D1 16.8 ± 1.39 C1 16.8 ± 1.39 D1 0.72 ± 0.14 D1 0.70 ± 0.04 D1 0.73 ± 0.04 D1

5th 17.2 ± 0.94 E1 17.4 ± 0.73 CD1 17.9 ± 0.61 DE1 0.78 ± 0.10 DE 0.76 ± 0.07 D1 0.75 ± 0.05 D1

6th 18.6 ± 0.80 E1 18.7 ± 0.86 D1 18.6 ± 0.56 E1 0.81 ± 0.08 E1 0.82 ± 0.13 E1 0.80 ± 0.13 E1

Resin C

1st 2.97 ± 0.28 A1 2.12 ± 0.12 A1 2.02 ± 0.15 A1 0.27 ± 0.07 A1 0.24 ± 0.04 A1 0.19 ± 0.02 A3

2nd 1.13 ± 0.16 B1 0.00 ± 0.00 B2 0.00 ± 0.00 B2 0.04 ± 0.01 B1 0.00 ± 0.02 B2 0.00 ± 0.00 B2

3rd 5.77 ± 1.21 C1 5.51 ± 0.96 C1 5.40 ± 1.55 C1 0.44 ± 0.09 A1 0.29 ± 0.06 A2 0.26 ± 0.06A A2

4th 15.0 ± 3.05 D1 11.4 ± 0.72D2 11.8 ± 1.54 D2 0.68 ± 0.05 C1 0.66 ± 0.04 C1 0.68 ± 0.05 C1

5th 16.8 ± 1.17 DE1 16.9 ± 1.52 E1 16.1 ± 1.00 E1 0.76 ± 0.05 D1 0.74 ± 0.07 D1 0.73 ± 0.04 C1

6th 17.9 ± 0.53 E1 17.8 ± 0.72 E1 18.1 ± 1.08 E1 0.81 ± 0.03 E1 0.81 ± 0.06 E1 0.76 ± 0.09 C1

Mean (S.D) of AFM nano-properties (Ei/Hi) of the dentine-bonded interfaces using the three different experimental resins used in this study

Same letter indicates no differences in columns (P [ 0.05) along different indentation position. Same number indicates no differences in rows for

different SBFS storage time (P [ 0.05); [Ei or Hi]

J Mater Sci: Mater Med (2012) 23:1521–1532 1525

123

therapeutic effects of remineralisation on the mineral-

depleted areas with the bonded-dentine interface must be

rejected.

Indeed, the results obtained during the AFM nano-

indentation assessment after prolonged SBFS storage

showed a statistical increase of the nano-properties (Ei and

Hi) in definite areas of the bonded-dentine interface created

using the two experimental resins containing BAG or BAG-

Zn micro-fillers. These results are in agreement with those

obtained by Bertassoni et al. [19] who showed that a con-

tinuous delivery of calcium (Ca2?) and phosphates (PO4-3)

ions may induce remineralisation and mechanical recovery

of mineral-deficient dentine due to a fine association of the

minerals with the organic matrix. Balooch et al. [20] dem-

onstrated that the AFM nano-indentation assessment, where

force and indenter displacement are recorded simulta-

neously during the indentation and the elastic modulus and

hardness are determined from the load displacement curve,

is a suitable method to evaluate the visco-elasticity of the

demineralised dentine and the effective remineralisation of

the collagen matrix [21, 22]. The results obtained in this

study confirmed these latter observations showing an HL

Fig. 2 Mean (S.D) of AFM nano-properties (Ei and Hi) of the main

zones along the resin–dentine interface created using the three

different experimental resins. The resin–dentine interface created

using resin A shows reduction of the adhesive layer’s nano-

mechanical properties and increase in the HL and bottom of HL

after SBFS storage (A and B). Resin B presents an increase in Hi

(C) and Ei (D) at the HL and bottom of HL after SBFS storage (1 m

and 3 m). While, an important nano-properties reduction can be

observed in the adhesive layer after prolonged SBFS storage. The

resin–dentine interface created with resin C is characterised by nano-

property reduction both at the HL and bottom of HL after SBFS

storage (E and F)

1526 J Mater Sci: Mater Med (2012) 23:1521–1532

123

with high visco-elasticity characteristics after 24 h of SBFS

storage and a well pronounced indentation mark only after

prolonged SBFS storage (Fig. 3C, 4A).

It is well known that BAG is able to bond hard and soft

tissues and evoke apatite precipitation due to a simultaneous

process where silicic acid Si(OH)4 is firstly released and

then chemisorbed within the demineralised collagen via

electrostatic bonding to form a poly-condensation silica-

gel-layer for the precipitation of Ca2? and PO4-3 [22–27].

Conversely, Zinc-doped bioactive glasses such as Bioglass

46S6 are characterised both by slower poly-condensation

reaction and Ca2? and PO4-3 release when immersed in

SBFS [28, 29]. Furthermore, the use of a polyacrylic acid

may act as a sequestrating agent and favour the aggregation

of Ca/P nano-precursors within the dentine collagen fibrils

regulating the entire process of remineralisation [33, 34].

The confocal micro-Raman spectroscopy results attained in

this study confirmed the slower apatite deposition activity

of the BAG-Zn-containing resin B which showed a con-

sistent presence of apatite only after prolonged (3 m) SBFS

storage. Conversely, the presence of the BAG micro-filler

induced a strong and consistent apatite precipitation

immediately after 1 m of SBFS storage (Fig. 1). Moreover,

the high zinc-content of the BAG-Zn (20 wt%) may have

protected the collagen fibrils from MMPs [13] which are

principally responsible for the degradation process both in

carious dentine [30–32] and within the HL [35, 36]. The

creation of a moderate alkaline environment within the

bonded-dentine interface due to the exchange between Na?

and H?/H3O? ions and the rapid release of Ca2? and PO4-3

species from the bioactive glasses may also have contrib-

uted in the inhibition of pH-dependent MMPs [13, 35, 36].

The confocal microscopy analysis offered further evi-

dence on the therapeutic ability of the two experimental

resins containing bioactive micro-fillers showing a strong

reflection signal and a reduced micropermeability along the

entire bonded-dentine interface after prolonged SBFS

storage (Figs. 3F and 4D). Conversely, a clear microper-

meability within a gap-affected interface was observed

within the dentine bonded with resin C (Fig. 5B). Similar

results were also observed in previous studies [37, 38]

which showed a progressive dehydration and water

replacement by mineral crystals within the denuded col-

lagen matrices of bonded-dentine interfaces.

Ion-releasing resin-based restorative materials, e.g.

resin-modified glass ionomer cements (RMGIC), are char-

acterised by an important water sorption phenomenon

which represents a key factor for their therapeutic activity

Fig. 3 (A) AFM topographic image showing the indentations per-

formed in a straight line from the adhesive layer (a) along the bonded-

dentine interface created with the resin A. The second indentation was

obtained from the HL (pointer) and the third value from the bottom of

the HL. The remnant indentations were performed along the intertu-

bular dentine (d). It is possible to see that the indentation mark

corresponding to the HL is scarcely visible (pointer). The appearance

of the first resin tags (t) after the bonding interface indicates the bottom

of the HL. The indentation mark performed on the HL of resin A stored

in SBFS for 1 m (B) and for 3 m (C) results well pronounced. (D) A

CLSM single projection image showing a bonded-dentine interface

(resin A) characterised by a clear HL, long resin tags penetrating the

dentinal tubules (t) and several filler lacunas (FL) within the adhesive

layer. In fluorescence mode it is possible to observe a porous HL

completely infiltrated by the fluorescent dye LY (pointer), (E) while,

in reflection/fluorescence mode it is possible to note a strong reflection

signal both from the dentinal tubules and from an HL only slightly

affected by LY micropermeability (arrow), (F)

J Mater Sci: Mater Med (2012) 23:1521–1532 1527

123

on demineralised dental tissues [39]. The resin blend solu-

tion used to formulate the resin bonding systems tested in

this study was specifically created (Table 1) to achieve an

essential resin elution/degradation during the SBFS storage

in order to favour ion-exchange and mineral precipitation

within the dentine-bonded interface [10, 11]. This is the

reason why the dentine-bonded specimens presented high

lTBS values after 24 h and an important lTBS reduction

(P \ 0.05) subsequent to 3 months of SBFS storage

(Table 2). Nevertheless, the SEM ultra-morphology analy-

sis performed in the 3 months SBFS-aged specimens cre-

ated with resin A and B showed a de-bonded surface

characterised by a remineralised resin-dentine matrix

(Fig. 6B2, C2). These results are in accordance with those

shown by Liu et al. [40] who identified large mineral plates

within a water-entrapped polymer hydrogel matrix corre-

sponding to the basal region of the filled water channels and

agglomerates of apatite nanocrystals within the gap region

of the water channel. However, it is also possible that the

crystallised SBFS-aged bonded-dentine interface created

using resin A and B may have had similar mechanical

characteristics as well as those created by GIC applied on

PAA-etched dentine when submitted to tensile tests. Indeed,

previous studies indicated that tensile or shear bond

strengths were rarely above 5 MPa [41, 42] while, the lTBS

of the GICs bonded to dentine was 12 ± 15 MPa with a

failure mode prevalently adhesive and mix [42]. Yip et al.

[43] stated that the results obtained from GICs bonded to

dentine under tensile stress does not represent their true

adhesive strength but only ultrastructural observation of the

bonding interface or of the fractured interface may indicate

the real bonding ability of conventional GIC to dentine.

However, further studies are ongoing on the formulation of

resin bonding systems containing bioactive micro-fillers

and biomimetic agents which may induce higher intra and

extra fibrils remineralisation and higher bond strength

results after prolonged SBFS aging.

In conclusion, a specific formulation of resin bond-

ing systems containing bioactive fillers such as BAG or

BAG-Zn may offer the possibility of increasing the

nano-mechanical properties of Hi and Ei and reduce the

micropermeability along the dentine-bonded interface by

therapeutic remineralisation of imperfect mineral-depleted

areas. This particular approach may be suitable for the

current concept of minimally invasive cavity preparation

followed by therapeutic restorative materials which may

stabilise the carious lesion and/or create an optimal envi-

ronment to repair the demineralised dental tissues.

Fig. 4 (A): AFM topographic

image showing the resin–

dentine interface created with

resin B after SBFS (3 m)

storage showing a clear

indentation mark on the HL.

(B): This CLSM single

projection captured from the

same resin-dentine specimens

(resin B) shows a clear HL with

long resin tags penetrating the

dentinal tubules (t) and a

substantial presence of BAG-Zn

filler (FL) within the adhesive

layer. (C): In reflection/

fluorescence lacune mode it is

possible to see a resin–dentine

interface characterised by a

porous HL completely

infiltrated by the LY dye

(pointer) after 24 h of SBFS

storage. (D): The CLSM single

projection image captured from

resin–dentine interfaces (resin

B) after prolonged SBFS

storage (3 m) showing a strong

reflection signal both in HL and

dentinal; a limited

micropermeability at the bottom

of the HL is also evident

(pointer)

1528 J Mater Sci: Mater Med (2012) 23:1521–1532

123

Fig. 5 (A): This CLSM reflection/fluorescence single projection

shows the resin–dentine interface created with resin C characterised

by a HL affected by severe micropermeability and a filler-free

adhesive layer (a). (B): This type of bonded-dentine interface showed

signs of degradation (pointer) between a remnant porous HL

infiltrated by LY dye after prolonged SBFS storage. No indentation

mark is visible both in the AFM image obtained in the specimens

stored for 24 h (C) and 3 m in SBFS (D). It is also possible to note the

presence of a gap between the adhesive layer (a) and the dentine

(d) indicating a severe degradation of the HL

Table 3 Mean and standard deviation of microtensile bond strength values in MPa obtained for the different experimental groups and per-

centage distribution of failure mode after microtensile bond strength testing; total number of beams (tested stick/pre-load failure)

lTBS-Mean ± SD (N of beams) % Failure [A/M/C] t P

24 h test 3 Month test

Resin A

(BAG)

B144.42 ± 7.43

(28/0) [15/70/15]

B215.02 ± 8.15

(25/3) [41/59/0]

11.66 \0.0001

Resin B

(BAG-Zn)

A136.60 ± 11.32

(30/0) [12/68/20]

A28.15 ± 5.12

(29/1) [72/26/2]

11.44 \0.0001

Resin C (No Filler) A137.44 ± 14.75

(31/0) [12/70/18]

C223.78 ± 10.11

(29/2) [32/68/0]

3.64 \0.001

For each horizontal row: values with identical numbers indicate no significant difference. P and t values are displayed at the right column

For each vertical column: values with identical letters indicate no significant difference using Student–Newman–Keuls test (P [ 0.05)

Failure mode [A adhesive, M mixed, C cohesive]

J Mater Sci: Mater Med (2012) 23:1521–1532 1529

123

Fig. 6 Scanning electron microscopy images of failure modes of

resin A (A1), resin B (A2) and resin C (A3) bonded to H3PO4-etched

dentine. Resin A (B1) shows a fracture surface after 3 m of SBFS

storage characterised by the presence of remnant resin (pointer) and

mineralised tissue (pointer). At higher magnification (B2) it was

possible to observe the presence of mineral crystals embedded in a

remnant resin/collagen scaffold-like structure (pointer). Conversely, a

limited presence of mineral crystals inside the tubules can only be

observed (pointer) in the specimens stored in SBFS for 24 h (B3).

Also the specimens created of the resin B group immersed in SBFS

for 3 m show remnant resin (pointer) on the fractured surface (C1).

At higher magnification it was possible to observe the presence of

mineral crystals (pointer) and no exposed collagen fibrils (C2). Also

in this case, the specimens aged in SBFS for 24 h show very few

mineral crystals embedded in the resin tags inside the tubules

(pointer) (C3). The SEM specimens of resin C group show a fractured

surface characterised by the presence of many exposed dentinal

tubules and remnant resin tags (D1). At higher magnification, it was

possible to note the presence of collapsed collagen fibrils (pointer),

exposed dentinal tubules (t) and remnant resin tags (rt) (C2). The 24 h

SBFS-aged bonded-dentine specimens showed a few exposed

dentinal tubules (pointer) and no exposed collagen fibrils (D3)

1530 J Mater Sci: Mater Med (2012) 23:1521–1532

123

Acknowledgments This work was supported by grants CICYT/

FEDER MAT2008-02347, CICYT/FEDER MAT2011-24551,

JA-P07-CTS2568 and JA-P08-CTS-3944. This article presents inde-

pendent research commissioned by the National Institute for Health

Research under the Comprehensive Biomedical Research Centre at

Guy’s & St. Thomas’ Trust. The views expressed in this publication

are those of the author(s) and not necessarily those of the NHS, the

NIHR or the Department of Health. The authors also acknowledge

support from the Centre of Excellence in Medical Engineering funded

by the Wellcome Trust. The authors have no financial affiliation or

involvement with any commercial organization with direct financial

interest in the materials discussed in this manuscript. Any other

potential conflict of interest is disclosed.

References

1. Tay FR, Pashley DH. Guided tissue remineralisation of partially

demineralised human dentine. Biomaterials. 2008;29:1127–37.

2. Liu Y, Mai S, Li N, Yiu CK, Mao J, Pashley DH, Tay FR.

Differences between top-down and bottom-up approaches in

mineralizing thick, partially demineralized collagen scaffolds.

Acta Biomater. 2011;7:1742–51.

3. Peters MC, Bresciani E, Barata TJ, Fagundes TC, Navarro RL,

Navarro MF, Dickens SH. In vivo dentin remineralization by

calcium-phosphate cement. J Dent Res. 2010;89:286–91.

4. Bresciani E, Wagner WC, Navarro MF, Dickens SH, Peters MC.

In vivo dentin microhardness beneath calcium-phosphate cement.

J Dent Res. 2010;89:836–41.

5. Moshaverinia A, Ansari S, Moshaverinia M, Schricker SR, Chee

WW. Ultrasonically set novel NVC-containing glass-ionomer

cements for applications in restorative dentistry. J Mater Sci

Mater Med. 2011;22:2029–34.

6. Dickens SH, Flaim GM. Effect of a bonding agent on in vitro

biochemical activities of remineralizing resin-based calcium

phosphate cements. Dent Mater. 2008;24:1273–80.

7. Dickens SH, Flaim GM, Takagi S. Mechanical properties and

biochemical activity of remineralizing resin-based Ca-PO4

cements. Dent Mater. 2003;19:558–66.

8. Ryou H, Niu LN, Dai L, Pucci CR, Arola DD, Pashley DH,

Tay FR. Effect of biomimetic remineralization on the dynamic

nanomechanical properties of dentin hybrid layers. J Dent Res.

2011;90:1122–8.

9. ten Cate JM, van Duinen RNB. Hypermineralization of dental

lesions adjacent to glass-ionomer cement restorations. J Dent Res.

1995;74:1266–71.

10. Tay FR, Pashley DH, Yiu C, Cheong C, Hashimoto M, Itou K,

Yoshiyama M, King NM. Nanoleakage types and potential

implications: evidence from unfilled and filled adhesives with the

same resin composition. Am J Dent. 2004;17:182–90.

11. Hashimoto M, Tay FR, Ohno H, Sano H, Kaga M, Yiu C,

Kumagai H, Kudou Y, Kubota M. Oguchi H.SEM and TEM

analysis of water degradation of human dentinal collagen. J Bio-

med Mater Res B Appl Biomater. 2003;66:287–98.

12. Pashley DH, Tay FR, Breschi L, Tjaderhane L, Carvalho RM,

Carrilho M, Tezvergil-Mutluay A. State of the art etch-and-rinse

adhesives. Dent Mater. 2011;7:1–16.

13. Osorio R, Yamauti M, Osorio E, Ruiz-Requena ME, Pashley DH,

Tay FR, Toledano M. Zinc reduces collagen degradation in

demineralized human dentin explants. J Dent. 2011;39:148–53.

14. Hench LL & Andersson O. Bioactive glasses. Introduction to

Bioceramics World Scientific, Singapore 1993; Hench LL, Wil-

son J eds: 45–47.

15. Edwards HGM, Carter, E.A. Biological Applications of Raman

Spectroscopy. Infrared and Raman Spectroscopy of Biological

Materials (Practical Spectroscopy) 2000; Gremlich, H.U., and

Yan, B. eds: 421–477.

16. Schulze KA, Oliveira SA, Wilson RS, Gansky SA, Marshall GW,

Marshall SJ. Effect of hydration variability on hybrid layer

properties of a self-etching versus an acid-etching system. Bio-

material. 2005;26:1011–8.

17. Sauro S, Osorio R, Watson TF, Toledano M. Assessment of

the quality of resin-dentin bonded interfaces: an AFM nano-

indentation, lTBS and confocal ultramorphology study. Dent

Mater. 2012. doi:10.1016/j.dental.2012.02.005.

18. Sauro S, Watson TF, Mannocci F, Miyake K, Huffman BP,

Tay FR, et al. Two-photon laser confocal microscopy of microp-

ermeability of resin-dentin bonds made with water or ethanol

wet bonding. J Biomed Mater Res B Appl Biomater. 2009;90:

327–37.

19. Bertassoni LE, Habelitz S, Kinney JH, Marshall SJ, Marshall GW

Jr. Biomechanical perspective on the remineralization of dentin.

Caries Res. 2009;43:70–7.

20. Balooch M, Wu-Magidi IC, Balazs A, Lundkvist AS, Marshall

SJ, Marshall GW, et al. Viscoelastic properties of demineralized

human dentin measured in water with atomic force microscope

(AFM)-based indentation. J Biomed Mater Res. 1998;40:539–44.

21. Balooch M, Habelitz S, Kinney J, Marshall S, Marshall G.

Mechanical properties of mineralized collagen fibrils as influ-

enced by demineralization. J Struct Biol. 2008;162:404–10.

22. Marshall GW Jr, Marshall SJ, Kinney JH, Balooch M. The dentin

substrate: structure and properties related to bonding. J Dent.

1997;25:441–58.

23. Engqvist H, Schultz-Walz JE, Loof J, Botton GA, Mayer D,

Phaneuf MW, et al. Chemical and biological integration of a

mouldable bioactive ceramic material capable of forming apatite

in vivo in teeth. Biomaterials. 2004;25:2781–7.

24. Hench LL. Bioceramics and the origin of life. J Biomed Mater

Res. 1989;23:685–93.

25. Zhong JP, LaTorre GP, Hench LL. The kinetics of bioactive

ceramics part VII: Binding of collagen to hydroxyapatite and

bioactive glass. In Bioceramics 7, eds. OH Andersson, A Yli-

Urpo, 1994, pp. 61–66.

26. Sauro S, Thompson I, Watson TF. Effects of common dental

materials used in preventive or operative dentistry on dentin

permeability and remineralization. Oper Dent. 2011;36:222–30.

27. Sepulveda P, Jones JR, Hench LL. Bioactive sol-gel foams for

tissue repair. J Biomed Mater Res. 2002;59:340–8.

28. Murphy S, Wren AW, Towler MR, Boyd D. The effect of ionic

dissolution products of Ca-Sr-Na-Zn-Si bioactive glass on in vitro

cytocompatibility. J Mater Sci Mater Med. 2010;21:2827–34.

29. Yamaguchi M, Igarashi A, Ychiyama S. Bioavailability of zinc

yeast in rats: stimulatory effect on bone calcification in vivo.

J Health Sci. 2004;50:75–81.

30. Tjaderhane L, Larjava H, Sorsa T, Uitto VJ, Larmas M, Salo T.

The activation and function of host matrix metalloproteinases in

dentin matrix breakdown in caries lesions. J Dent Res. 1998;

77:1622–9.

31. van Strijp AJ, Jansen DC, DeGroot J, ten Cate JM, Everts V.

Host-derived proteinases and degradation of dentine collagen in

situ. Caries Res. 2003;37:58–65.

32. Chaussain-Miller C, Fioretti F, Goldberg M, Menashi S. The role

of matrix metalloproteinases (MMPs) in human caries. J Dent

Res. 2006;85:22–32.

33. Liu Y, Li N, Qi YP, Dai L, Bryan TE, Mao J, Pashley DH, Tay

FR. Intrafibrillar collagen mineralization produced by biomimetic

hierarchical nanoapatite assembly. Adv Mater. 2011;23:975–80.

34. Liu Y, Tjaderhane L, Breschi L, Mazzoni A, Li N, Mao J,

Pashley DH, Tay FR. Limitations in bonding to dentin and

experimental strategies to prevent bond degradation. J Dent Res.

2011;90:953–68.

J Mater Sci: Mater Med (2012) 23:1521–1532 1531

123

35. Mazzoni A, Pashley DH, Nishitani Y, Breschi L, Mannello F,

Tjaderhane L, et al. Reactivation of inactivated endogenous

proteolytic activities in phosphoric acid-etched dentine by etch-

and-rinse adhesives. Biomaterials. 2006;27:4470–6.

36. De Munck J, Mine A, Van den Steen PE, Van Landuyt KL,

Poitevin A, Opdenakker G, Van Meerbeek B. Enzymatic degra-

dation of adhesive-dentin interfaces produced by mild self-etch

adhesives. Eur J Oral Sci. 2010;118:494–501.

37. Kim J, Gu L, Breschi L, Tjaderhane L, Choi KK, Pashley DH,

Tay FR. Implication of ethanol wet-bonding in hybrid layer

remineralization. J Dent Res. 2010;89:575–80.

38. Kim J, Vaughn RM, Gu L, Rockman RA, Arola DD, Schafer TE,

et al. Imperfect hybrid layers created by an aggressive one-step

self-etch adhesive in primary dentin are amendable to biomimetic

remineralization in vitro. J Biomed Mater Res A. 2010;93:

1225–34.

39. Toledano M, Osorio R, Osorio E, Aguilera FS, Romeo A, de la

Higuera B. Garcıaet al. Sorption and solubility testing of ortho-

dontic bonding cements in different solutions. J Biomed Mater

Res B Appl Biomater. 2006;76:251–6.

40. Liu G, Zhao D, Tomsia AP, Minor AM, Song X, Saiz E. Three-

dimensional biomimetic mineralization of dense hydrogel tem-

plates. J Am Chem Soc. 2009;29:9937–9.

41. Hewlett ER, Caputo AA, Wrobet DC. Glass ionomer bond

strength and treatment of dentin with polyacrylic acid. J Prosthet

Dent. 1991;66:767–72.

42. Berry EA, Powers JM. Bond strength of glass ionomers to

coronal and radicular dentin. Oper Dent. 1994;19:122–6.

43. Yip HK, Tay FR, Ngo HC, Smales RG, Pashley DH. Bonding of

contemporary glass ionomer cements to dentin. Dent Mater.

2001;17:456–70.

1532 J Mater Sci: Mater Med (2012) 23:1521–1532

123


Recommended