Journal List > J Adv Prosthodont > v.3(1) > 1054010

Sohn, Heo, Koak, Kim, and Lee: Strain of implants depending on occlusion types in mandibular implant-supported fixed prostheses

Abstract

PURPOSE

This study investigated the strain of implants using a chewing simulator with strain gauges in mandibular implant-supported fixed prostheses under various dynamic loads.

MATERIALS AND METHODS

Three implant-supported 5-unit fixed prostheses were fabricated with three different occlusion types (Group I: Canine protected occlusion, Group II: Unilaterally balanced occlusion, Group III: Bilaterally balanced occlusion). Two strain gauges were attached to each implant abutment. The programmed dynamic loads (0 - 300 N) were applied using a chewing simulator (MTS 858 Mini Bionix II systems, MTS systems corp., Minn, USA) and the strains were monitored. The statistical analyses were performed using the paired t-test and the ANOVA.

RESULTS

The mean strain values (MSV) for the working sides were 151.83 µε, 176.23 µε, and 131.07 µε for Group I, Group II, and Group III, respectively. There was a significant difference between Group II and Group III (P < .05). Also, the MSV for non-working side were 58.29 µε, 72.64 µε, and 98.93 µε for Group I, Group II, and Group III, respectively. One was significantly different from the others with a 95% confidence interval (P < .05).

CONCLUSION

The MSV for the working side of Groups I and II were significantly different from that for the non-working side (Group I: t = 7.58, Group II: t = 6.25). The MSV for the working side of Group II showed significantly larger than that of Group III (P < .01). Lastly, the MSV for the non-working side of Group III showed significantly larger than those of Group I or Group II (P < .01).

INTRODUCTION

Dental implants were introduced in the late 1960s for the rehabilitation of the completely edentulous patients.1,2 Through the preliminary studies on osseointegration, dental implants have been extensively used for the rehabilitation of completely or partially edentulous patients over the last three decades.1,3 And the use of implants has revolutionized dental treatment modalities and provided excellent long-term results.4 Despite high success rates reported by a vast number of clinical studies, early or late implant failures are still unavoidable.5 Late implant failures are observed after prosthesis delivery and are mainly related to biomechanical complications. One etiologic factor is the lateral component of force during the transmission of forces by prosthesis to dental implants. The lateral component of force is responsible for creating bending moments which have a destructive effect on the cortical bone surrounding the implant collar and can cause complications associated with component loosening or fracture.5 Yet, the mechanisms responsible for biomechanical implant failures are not fully understood and the literature concerning the influences of several biomechanical factors are inconclusive.6
Occlusion can be critical for implant longevity because of the nature of the potential load created by tooth contacts and the impact on the attachment of the bone to the titanium implant. In the natural dentition, the periodontal ligament has the capacity to absorb stress or allow for tooth movement, but the bone-implant interface seemingly has no capacity to allow movement of the implant.7,8
Vertical loads from mastication induce axial forces and bending moment and result in stress gradients in the implant as well as in the bone. A key factor for the success or failure of a dental implant is the manner in which stress is transferred to the surrounding bone.9
Deflective contacts in the intercuspal position may be responsible for excessive force development. Gibbs et al.10 reported that the greatest forces during mastication are exerted in the intercuspal position. If this position is unstable, intolerably high forces can be exerted. The application of functional forces induces stress and strain within the implant-prosthesis complex and affects the bone remodeling process around implants.2,11 Yet, the physiologic tolerance thresholds of human jawbones are not known and some reported implant failures may be related to magnitudes of stress beyond tolerable levels.
Studies on bone biology suggest that implant overloading may lead to implant failures when implants are overloaded, high deformations (above 2000 - 3000 microstrain) occur in the bones surrounding the implants.12 When pathologic overloading occurs (over 4000 microstrain), stress and strain gradients exceed the physiologic tolerance threshold of the bones and cause microfractures at the bone-implant interface.13 While overloading may be manifested by the application of repeated single loads, which causes micro-fractures within the bone tissue, continuous application of low loads may also lead to failure, namely, fatigue fracture. Excessive dynamic loading may also decrease bone density around the neck of implants and lead to crater-like defects.14 Accordingly, overload-associated implant failures have been reported following the first year of prosthodontic treatment.15
Horshaw and co-workers16 reported that overloading of implants resulted in an increased bone resorption around the implant collar, and a decreased percentage of mineralized bone tissue in the cortex within 350 um of the implant was evident after 12 weeks of load application. Marginal bone resorption may also be related to the lack of mechanical coupling between the machined coronal region of the implant and the bone, which prevents effective transfer of occlusal forces from the implant to the cortical bone. The extremely low intraosseous strain ( ≥ below 100 microstrain) thus causes bone resorption due to disuse atrophy.17,18 If the surface of the implant is rough, the total area used to transfer occlusal forces to the bone increases. Eventually, lower stress and strain can be achieved in the vicinity of the implant. Rough surface implants also provide better mechanical interlock with the bone than machined-surface implants do.19
The type of prosthesis affects the mode of implant loading. In cement-retained implant restorations, the occlusal surface is devoid of screw holes and the occlusion can be developed that responds to the need for axial loading. Screwretained fixed prosthesis or overdentures, however, are subjected to off-set loads that cause a substantial increase in bending moments.20
Mericske-Stern and collaborators23 also registered forces on implants supporting one-piece full-arch fixed prosthesis and bar-retained overdentures in the maxilla. They concluded that, the type of prosthesis did not have a determining effect on the force pattern. However, in overdenture treatments, the resorption pattern of the maxilla affects positioning of the implants and the denture teeth. Since the positioning of denture teeth frequently creates an anterior or labial cantilever, which acts as a long lever-arm, high bending moments are created on maxillary implants. This situation may explain why implant survival rates are significantly lower in the maxilla, particularly with overdenture treatments.22,23
Regardless of the design, the occlusal materials for implant-prosthesis complexes have been a topic of research interest in recent studies. Skalak24 envisaged that the use of acrylic resin teeth would be useful for shock protection on implants and Brånemark and co-workers2 have also recommended the use of acrylic resin as the material of choice for the occlusal surfaces of implant-retained prostheses. The resiliency of this material was suggested as a safeguard against the negative effects of impact forces and microfractures at the bone-implant interface. The literature, however, is inconclusive on its effect on shock absorption.25-29 In fact, acrylic resins seem to prevent technical problems and subjective disadvantages. For example, due to their low wear resistances, premature contacts often occur after several months of prosthesis delivery. On the other hand, gold and porcelain surfaces are believed not to provide force absorption, but they are also frequently used. Although the choice of prosthesis material still remains as a topic of controversy and argument, there is a consensus that it does not have any influence on implant survival.30
When dealing with a complex stress analysis problem in which a complete theoretical solution may prove impractical with respect to time, cost, or degree of difficulty, experimental techniques are often used. Current techniques employed to evaluate the biomechanical loads on implants comprise the use of mathematical calculations, photoelastic stress analysis, two or three-dimensional finite element stress analysis and strain-gauge analysis (SGA). The application of SGA on dental implants is based on the use of electrical resistance strain-gauges and its associated equipment, and provides both in-vivo and in-vitro measurement of strain under static or dynamic loads. Under an applied force, a strain gauge measures the mean dimensional change where it is bonded or embedded. For in-vivo or in-vitro strain-gauge experimentation, however, this may not be provided due to several factors included in force transmission during load application by the opposing teeth or by an apparatus. The placement of the gauges may have slight inaccuracies, or the angulation of implants may not be as precise as in a theoretical model. Overall, the very nature of the physical experimental technique makes it inherently subject to random errors. Currently, although SGA is the only technique that allows in-vivo measurements during clinical loading, the results of in-vivo and in-vitro SGA do not agree on the quantification of bending moments.31,32
Correct qualification and quantification of forces on implants are crucial for understanding the biomechanics of implants. Biomechanical studies should, therefore, be designed not only for descriptive purposes but also for obtaining reliable and accurate data that have clinical relevance.
The purpose of this study is to investigate the strain of implants using chewing simulator (MTS 858 Mini Bionix II systems, MTS systems corp., Minn, USA) combined with strain gauge in mandibular implant-supported fixed prostheses under various dynamic loads and to evaluate and compare the quantity and distribution of mean strain values on the working and non-working sides of implant-supported fixed prostheses.

MATERIALS AND METHODS

1. Fabrication of measurement model

A partially edentulous mandibular acrylic resin model and opposing teeth were fabricated with auto-polymerizing resin (POLYUROCK; Metalor technologies, Stuttgart, Swiss) and artificial denture teeth (Endura; Shofu inc., Kyoto, Japan). This auto-polymerizing resin has a flexural modulus of 3,000 MPa; similar to that of the mandibular trabecular bone.33 The areas from the mandibular canine to the second molar were specially designed to be the missing span for the installation of implants and screw-retained fixed prostheses.
The interarch distance from the alveolar ridge of mandibular missing span to the cusp of opposing teeth was endowed at least with 12 mm-apart, because strain gauges and the abutment complex need sufficient vertical space (> 5 mm). After fabrication of implant surgical stent, six dental implants, ф4.0 mm × 11.5 mm height (US II; Osstem, Seoul, Korea), were installed at the mandibular canine (Fx1, Fx4), the second premolar (Fx2, Fx5) and the second molar (Fx3, Fx6) areas in the missing spans by using standard surgical procedures in real clinical situations. And then, six dental implants were endowed with 30 N torque value of initial stability by using an electric torque measuring device. All extended vertical axes of the six implants were headed for functional cusps of the opposing teeth, so that screw holes of screw-retained implant prostheses could be positioned centrally.

2. Fabrication of implant-supported fixed prostheses

Three implant-supported 5-unit screw-retained fixed prostheses were made by using standard methods of superstructure fabrication. The entire procedure, including impression taking, master cast fabrication, wax-up, casting, and finishing, was carried out in accordance with recommended protocols.34,35 The fixture-level impression was taken using an individual resin tray and pick-up type impression copings with polyvinyl-siloxane impression material (Imprint II; 3M ESPE, MN., USA).
The master cast was fabricated with stone and implant lab analogues. And then, standard abutments (ф4.5 mm × 5.5 mm height, Osstem, Seoul, Korea) were connected to lab analogues. The mandibular cast and the opposing maxillary cast were mounted arbitrarily on a semi-adjustable articulator. The condylar and anterior guidances of the articulator were set at 30 degrees and the Bennett angle was set at 15 degrees.
After the wax-up procedure, gold frameworks were cast with Type III gold alloy (Harmony C&B55; Ivoclar-vivadent, Liechtenstein, Germany) and recommended protocols were performed to obtain a passive fit of the superstructures.
Reinforced hard resin (Tescera; Ivoclar-vivadent, Liechtenstein, Germany) was selected as the material for the occlusal table because it is favorable to transfer force and easy to add or eliminate, allowing application of different occlusion types. The occlusal table was designed to have more than three occlusal contacts on the functional cusp, central fossa and marginal ridge of each tooth.

3. Endowment with three different occlusion types

To minimize errors, by addition or elimination of Tescera materials, three test groups of implant-supported 5-unit fixed prostheses with different occlusion types were fabricated on the same supra-structure. The following are three different groups with different occlusion types of implant-supported fixed prostheses.
  • Group I: Canine protected occlusion

  • Group II: Unilaterally balanced occlusion

  • Group III: Bilaterally balanced occlusion

4. Attachment of strain gauges and equipment set-up

The surfaces of six standard abutments were roughened by sandblasting media and marked with a felt-tipped pen bucco-lingually. Two strain gauges (KFG-1-120-C1-11L1M2R; KYOWA electronic instruments, Tokyo, Japan) were then attached with adhesive (M-bond 200; Tokuyama, Tokyo, Japan) bucco-lingually on each standard abutment (Fig. 1). A total of twelve strain gauges (Ch1 - Ch6 in working side, the other six Ch7 - Ch12 in non-working side) were attached on test model and connected to dynamic signal conditioning strain amplifiers (CTA1000, Curiotech inc., Paju, Korea). The amplified dynamic strain signals were monitored with a software program (DA1700, National Instrument., TX, USA).

5. Control loading and monitoring dynamic strain signals

The mandibular test models with different occlusion types and the opposing teeth models were mounted on a MTS chewing simulator (MTS 858 Mini Bionix II systems, MTS systems corp., MN, USA) using a custom-made positioning jig. Then, six channels of strain gauges on the mandibular working side were connected to dynamic signal conditioning strain amplifiers and a zero point calibration was performed (Fig. 2). The programmed vertical and lateral combined dynamic loads, 0 - 300 N in 25 N increments, were applied on the mandibular test model twenty-five times, and ten stable dynamic strain signals were monitored and recorded. After monitoring the signals on the working side, the same experiments were repeated for the other six channels on the non-working side. Figure 3 shows the profile curve of the dynamic load controlled by MTS chewing simulation function. The blue line represents the quantity and distribution of the vertical axial load, and the red line those of the concurrent lateral shear strength.

6. Statistical analysis

First, strain values of the working side of each group under different dynamic loads were compared with those of the non-working side using the paired sample t-test. Second, multiple comparisons of three implants positioned on the working side in each group were performed using the one-way ANOVA with the Tukey HSD (Honestly Significant Difference) multiple comparison method. Last, multiple comparisons of the results from the working and non-working sides in the three groups were also performed using the same methods.

RESULTS

1. The mean strain value of the implants in different groups

An example of curves of strain on working side of Group I under 100 N dynamic load were drawn in complex shapes, as in figure 4. The strain values of the buccally positioned Ch1, 3, and 5 strain gauges of the working side and those of the lingually positioned Ch8, 10 and 12 strain gauges in each group represent (+) tensile strain values. On the other hand, the values of the lingually positioned Ch2, 4 and 6 strain gauges of the working side and those of the buccally positioned Ch7, 9, and 11 strain gauges in each group represent (-) compressive strain values.
The arithmetical absolute mean value of two strain gauges attached to each implant fixture was calculated. Table 1, 2, and 3 show the mean absolute strain value of each implant fixture, and Figure 5, 6 and 7 show the curves of strain values under increasing dynamic loads from 25 N to 300 N.

2. Comparative analysis of the working side with the non-working side

In order to compare the mean strain values of the working side with those of the non-working side in each group, paired sample t-tests were carried out. In Group I, there was a significant difference between the mean strain value of working side and that of the non-working side with a t-value of 7.58. In Group II, the same was observed with a t-value of 6.25. In Group III, however, a much smaller t-value (3.83) was observed, showing a difference not as significant as observed in Group I or Group II (Table 4).

3. Multiple comparisons of three implants positioned on the working side

In order to compare the mean strain values of the three implants positioned on the working side (Fx1, 2, and 3) of each group statistically, multiple comparisons were made using the Tukey's HSD (Honestly Significant Difference) methods (Table 5). In the case of Group I, the three P values (significance probabilities) between the groups are below .05, so any one of the three mean strain values may be significantly different from the others, with a 95% confidence interval. And the P-value of Fx2 - Fx3 is 0.039, which is higher than the others (P < .01), it means that the statistical difference is comparatively lower than those of Fx1 - Fx2 or Fx1 - Fx3. In the case of Group II, the same trend was observed as in Group I: the three P values of between the groups are below 0.01.
In the last case of Group III, the P-values of Fx1 - Fx2 and Fx2 - Fx3 are below .05. But the P-value of Fx1 - Fx3 is 0.103 (P > .05). Therefore, the mean strain value of Fx1 was not different from that of Fx3 with a 95% confidence interval.

4. Multiple comparisons of the mean strain values for the working or non-working side

At last, in order to compare the generalized mean strain values for the working or non-working side of the three groups, multiple comparisons were made using the Tukey's HSD methods, too.
The generalized mean strain values for the working side of Group I, Group II, and Group III were 151.83 µε, 176.23 µε, and 131.07 µε, respectively. There was no significant difference between Group I and Group II or between Group I and Group III with a 95% confidence interval (P > .05). But there was a significant difference between Group II and III (P < .05, Table 6).
The generalized mean strain values of non-working side of Group I, Group II, and Group III were 58.29 µε, 72.64 µε, and 98.93 µε, respectively. There were significant differences between any two among the three groups (P < .05, Table 6).

DISCUSSION

Curtis et al.36 reported that partial edentulous mandibles were more common than partial edentulous maxillae and the class I mandibular RPDs were the most common type for either dental arch and the percentage of Kennedy class I RPDs were 40%, class II 33%, class III 18%, and class IV were 9%. In this study, the reason for choosing the mandibular class I partial edentulous test model was that we had often been faced with this clinical situation. When anterior fixed ceramic restoration and bilateral posterior implant-supported fixed prostheses are combined, we should take occlusion-related factors into special considerations because the treatment would be mandibular full-mouth rehabilitation.
This study have realized in-vitro chewing simulations using the dynamic load control of the MTS machine and, it was the latest mechanical study of the implant-abutment complex on strain under various dynamic loads using the combined method of the MTS chewing simulation and strain gauge. So this technique will be a good methodological way for any other study of implant on strain under various dynamic loads containing lateral forces.
The methodological key point of this study was to produce three occlusion types on implant-supported fixed prostheses with inclined cusp angle and anterior guidance with predetermined mandibular condylar guidance. But the MTS machine was able to reproduce sufficiently the three dimensional movements of the real mandible.
For dentate humans, the maximum biting force varies among individuals and in different regions of the dental arch.37,38 The greatest maximum biting force reported to date is 443 kgN.39 Dentate patients have 5 - 6 times higher bite force than complete denture wearers.40 Present evidence based principally on static force measurements indicates that the average biting force is 100 - 150 N in adult males, and males have a higher biting force than females.37 Patients with implant-supported fixed prosthesis have a masticatory muscle function equal to or approaching that of patients with natural teeth or with tooth-supported fixed partial dentures.41 For this reason, the dynamic load of the MTS machine was set in the range of +25 N to +300 N in 25 N increments to include light chewing forces to maximal. A careful notice of position of strain gauge and data analysis was taken not to exaggerate error. To minimize errors in data analysis generated by incorrect position of the strain gauge, two strain gauges were attached to the abutment on both the buccal and the lingual side, and then the arithmetical absolute mean value was calculated.
The mean strain values of each group under the same mechanical conditions, except for the occlusion type, had different distributions and quantities depending on the implant position and the magnitude of the dynamic load. These data were analyzed by statistical methods to yield the following results.
First, the paired sample t-tests on the working side versus the non-working side of each group showed that the t-values of Group I and II were approximately twice that of Group III. This proved statistically that the mean strain values of the working side were significantly different from those of the non-working side in Group I and Group II but not in Group III. In other words, only bilaterally balanced occlusion type dispersed the occlusal loads evenly and a comparatively smaller strain was generated under the same dynamic load.
Multiple comparisons of the mean strain value for three implants positioned on the working side of Group I showed that the mean strain values were 220.15 µε, 136.09 µε, and 99.25 µε for Fx1, Fx2, and Fx3, respectively, and proved that any one was significantly different from the others with a 95% confidence interval (P < .05). This suggested that strain was mainly concentrated on an area of the working side canine and much more stress would be generated around the Fx1-abutment complex.
In Group II, multiple comparisons showed that the mean strain values were 241.97 µε, 179.87 µε, and 106.86 µε for Fx2, Fx3, and Fx1, respectively, and proved that any one was significantly different from the others with a 95% confidence interval (P < .05). This suggested that strain was mainly concentrated on an area of working side second premolar and much more stress would be generated around the Fx2-abutment complex. Mericske-Stern and Zarb42 investigated occlusal forces in a group of partially edentulous patients restored with ITIw implants supporting fixed partial prostheses and measured an average value of maximum occlusal force lower than 200 N in the first premolars and molars and 300 N in the second premolars. These data suggest that forces on implants are dependent on the location of the implant in the dental arch and implants placed in the posterior region of the mouth are at greater risk of overloading.
In Group III, multiple comparisons showed that the mean strain values were 165.48 µε, 120.42 µε, and 107.32 µε for Fx2, Fx3, and Fx1, respectively, and proved that there was no significant difference between Fx1 and Fx3 with a 95% confidence interval (P > .05). Unexpectedly, however, the mean strain value of Fx2 was significantly different from the others (P < .05). But, it is difficult to give this finding some significance because bilaterally balanced occlusion type disperses strain to all implants evenly.
Multiple comparisons of the generalized mean strain values for the working sides of the three groups showed that the mean strain values were 151.83 µε, 176.23 µε, and 131.07 µε for Group I, Group II, and Group III, respectively. There was a significant difference between Group II and Group III (P < .05), but the mean strain value for the working side of Group I was not significantly different from the others with a 95% confidence interval (P > .05). This means that only unilaterally balanced occlusion has more remarkable strain on the working side than the bilaterally balanced occlusion.
Lastly, multiple comparisons of the generalized mean strain values for non-working side of the three groups showed that the mean strain values were 58.29 µε, 72.64 µε, and 98.93 µε for Group I, Group II, and Group III, respectively. That of bilaterally balanced occlusion was significantly different from the others with a 95% confidence interval (P < .05).
These statistical analyses suggest that the strain of implants under dynamic load was related to the position of the implants in the dental arch and also to the occlusion type including the cusp inclination, the condylar guidance, and the anterior guidance.

CONCLUSION

In this study, strain gauge analyses using the MTS machine were carried out to evaluate the quantity and distribution of strain generated from three implant supported 5-unit fixed implant prostheses under dynamic load. The following conclusions were drawn.
  1. The mean strain values for the working side of Groups I and II were significantly different from that for the non-working side (t = 6.2 - 7.5). But, Group III (t = 3.8) was not.

  2. The comparative sequence of the mean strain values for the three implants of Group I working side was Fx1 > Fx2 > Fx3; one was significantly different from the others (P < .05).

  3. The comparative sequence of the mean strain values for the three implants of Group II working side was Fx2 > Fx3 > Fx1; one was significantly different from the others (P < .01).

  4. The comparative sequence of the mean strain values for the three implants of Group III working side was Fx2 > Fx3 ≒ Fx1; there was no significant difference between Fx1 and Fx3 (P > .05).

  5. Multiple comparisons of the generalized mean strain values for the working side of the three groups showed that only the value of Group II was significantly larger than that of Group III (P < .01).

  6. Multiple comparisons of the generalized mean strain values for the non-working side of the three groups showed that the value of Group III was significantly larger than those of Group I or Group II (P < .01).

Figures and Tables

Fig. 1
Graphic illustration of the measurement model and strain gauges.
jap-3-1-g001
Fig. 2
The overview of test model mounted on MTS machine.
jap-3-1-g002
Fig. 3
The controlled profile curve of dynamic load.
jap-3-1-g003
Fig. 4
An example of curves of strain on working side of Group I under 100 N dynamic load.
jap-3-1-g004
Fig. 5
The mean absolute strain value for the six implants of Group I.
jap-3-1-g005
Fig. 6
The mean absolute strain value for the six implants of Group II.
jap-3-1-g006
Fig. 7
The mean absolute strain value for the six implants of Group III.
jap-3-1-g007
Table 1
The mean absolute strain value for the six implants of Group I
jap-3-1-i001
Table 2
The mean absolute strain value for the six implants of Group II
jap-3-1-i002
Table 3
The mean absolute strain value for the six implants of Group III
jap-3-1-i003
Table 4
The paired sample t-tests of working side Vs non-working side in each group
jap-3-1-i004
Table 5
The multiple comparisons of the three implants positioned on working side
jap-3-1-i005

Tukey HSD (Honestly Significant Difference), *The P value is significant at the 0.05 level, **The P value is significant at the 0.01 level.

Table 6
The multiple comparisons of the mean strain values for the working or non-working side
jap-3-1-i006

Tukey HSD (Honestly Significant Difference), *The P value is significant at the 0.05 level, **The P value is significant at the 0.01 level.

References

1. Adell R, Eriksson B, Lekholm U, Brånemark PI, Jemt T. Long-term follow-up study of osseointegrated implants in the treatment of totally edentulous jaws. Int J Oral Maxillofac Implants. 1990. 5:347–359.
2. Brånemark PI, Zarb GA, Albrektsson T, editors. Tissue Integrated Prostheses: osseointegration in clinical dentistry. 1985. Chicago: Quintessence;175–186.
3. Adell R, Lekholm U, Rockler B, Brånemark PI. A 15-year study of osseointegrated implants in the treatment of the edentulous jaw. Int J Oral Surg. 1981. 10:387–416.
4. Esposito M, Hirsch JM, Lekholm U, Thomsen P. Biological factors contributing to failures of osseointegrated oral implants. (II). Etiopathogenesis. Eur J Oral Sci. 1998. 106:721–764.
5. Kohavi D. Complications in the tissue integrated prostheses components: clinical and mechanical evaluation. J Oral Rehabil. 1993. 20:413–422.
6. Taylor TD, Agar JR, Vogiatzi T. Implant prosthodontics: current perspective and future directions. Int J Oral Maxillofac Implants. 2000. 15:66–75.
7. Chapman RJ. Principles of occlusion for implant prostheses: guidelines for position, timing, and force of occlusal contacts. Quintessence Int. 1989. 20:473–480.
8. Ashman RB, Van Buskirk WC. The elastic properties of a human mandible. Adv Dent Res. 1987. 1:64–67.
9. Geng JP, Tan KB, Liu GR. Application of finite element analysis in implant dentistry: a review of the literature. J Prosthet Dent. 2001. 85:585–598.
10. Gibbs CH, Mahan PE, Lundeen HC, Brehnan K, Walsh EK, Holbrook WB. Occlusal forces during chewing and swallowing as measured by sound transmission. J Prosthet Dent. 1981. 46:443–449.
11. Bidez MW, Misch CE. Force transfer in implant dentistry: basic concepts and principles. J Oral Implantol. 1992. 18:264–274.
12. Stanford CM, Brand RA. Toward an understanding of implant occlusion and strain adaptive bone modeling and remodeling. J Prosthet Dent. 1999. 81:553–561.
13. Roberts WE. Naert I, van Steenberghe D, Worthington P, editors. Fundamental principles of bone physiology, metabolism and loading. Osseointegration in oral rehabilitation: an introductory textbook. 1993. London: Quintessence;163–164.
14. Duyck J, Rønold HJ, Van Oosterwyck H, Naert I, Vander Sloten J, Ellingsen JE. The influence of static and dynamic loading on marginal bone reactions around osseointegrated implants: an animal experimental study. Clin Oral Implants Res. 2001. 12:207–218.
15. Quirynen M, Naert I, van Steenberghe D. Fixture design and overload influence marginal bone loss and fixture success in the Brånemark system. Clin Oral Implants Res. 1992. 3:104–111.
16. Horshaw SJ, Brunski JB, Cochran GV. Mechanical loading of Brånemark implants affects interfacial bone modelling and remodelling. Int J Oral Maxillofac Implants. 1994. 9:345–360.
17. Frost HM. Wolff's Law and bone's structural adaptations to mechanical usage: an overview for clinicians. Angle Orthod. 1994. 64:175–188.
18. Pilliar RM, Deporter DA, Watson PA, Valiquette N. Dental implant design-effect on bone remodeling. J Biomed Mater Res. 1991. 25:467–483.
19. Buser D, Nydegger T, Hirt HP, Cochran DL, Nolte LP. Removal torque values of titanium implants in the maxilla of miniature pigs. Int J Oral Maxillofac Implants. 1998. 13:611–619.
20. Hebel KS, Gajjar RC. Cement-retained versus screw-retained implant restorations: achieving optimal occlusion and esthetics in implant dentistry. J Prosthet Dent. 1997. 77:28–35.
21. Mericske-Stern R, Venetz E, Fahrländer F, Bürgin W. In vivo force measurements on maxillary implants supporting a fixed prosthesis or an overdenture: a pilot study. J Prosthet Dent. 2000. 84:535–547.
22. Bergendal T, Engquist B. Implant-supported overdentures: a longitudinal prospective study. Int J Oral Maxillofac Implants. 1998. 13:253–262.
23. Lekholm U, Gunne J, Henry P, Higuchi K, Lindén U, Bergström C, van Steenberghe D. Survival of the Brånemark implant in partially edentulous jaws: a 10-year prospective multicenter study. Int J Oral Maxillofac Implants. 1999. 14:639–645.
24. Skalak R. Biomechanical considerations in osseointegrated prostheses. J Prosthet Dent. 1983. 49:843–848.
25. Davis DM, Rimrott R, Zarb GA. Studies on frameworks for osseointegrated prostheses: Part 2. The effect of adding acrylic resin or porcelain to form the occlusal superstructure. Int J Oral Maxillofac Implants. 1988. 3:275–280.
26. Gracis SE, Nicholls JI, Chalupnik JD, Yuodelis RA. Shock absorbing behavior of five restorative materials used on implants. Int J Prosthodont. 1991. 4:282–291.
27. Cibirka RM, Razzoog ME, Lang BR, Stohler CS. Determining the force absorption quotient for restorative materials used in implant occlusal surfaces. J Prosthet Dent. 1992. 67:361–364.
28. Hobkirk JA, Psarros KJ. The influence of occlusal surface material on peak masticatory forces using osseointegrated implant-supported prostheses. Int J Oral Maxillofac Implants. 1992. 7:345–352.
29. Sertgöz A. Finite element analysis study of the effect of superstructure material on stress distribution in an implant-supported fixed prosthesis. Int J Prosthodont. 1997. 10:19–27.
30. Brunski JB, Puleo DA, Nanci A. Biomaterials and biomechanics of oral and maxillofacial implants: current status and future developments. Int J Oral Maxillofac Implants. 2000. 15:15–46.
31. Glantz PO, Rangert B, Svensson A, Stafford GD, Arnvidarson B, Randow K, Lindén U, Hultén J. On clinical loading of osseointegrated implants. A methodological and clinical study. Clin Oral Implants Res. 1993. 4:99–105.
32. Smedberg JI, Nilner K, Rangert B, Svensson SA, Glantz SA. On the influence of superstructure connection on implant preload: a methodological and clinical study. Clin Oral Implants Res. 1996. 7:55–63.
33. Keaveny TM, Guo XE, Wachtel EF, McMahon TA, Hayes WC. Trabecular bone exhibits fully linear elastic behavior and yields at low strains. J Biomech. 1994. 27:1127–1136.
34. Heckmann SM, Karl M, Wichmann MG, Winter W, Graef F, Taylor TD. Cement fixation and screw retention: parameters of passive fit. An in vitro study of three-unit implant-supported fixed partial dentures. Clin Oral Implants Res. 2004. 15:466–473.
35. Karl M, Winter W, Taylor TD, Heckmann SM. In vitro study on passive fit in implant-supported 5-unit fixed partial dentures. Int J Oral Maxillofac Implants. 2004. 19:30–37.
36. Curtis DA, Curtis TA, Wagnild GW, Finzen FC. Incidence of various classes of removable partial dentures. J Prosthet Dent. 1992. 67:664–667.
37. Helkimo E, Carlsson GE, Helkimo M. Bite force and state of dentition. Acta Odontol Scand. 1977. 35:297–303.
38. van Eijden TM. Three-dimensional analyses of human bite-force magnitude and moment. Arch Oral Biol. 1991. 36:535–539.
39. Gibbs CH, Mahan PE, Mauderli A, Lundeen HC, Walsh EK. Limits of human bite strength. J Prosthet Dent. 1986. 56:226–229.
40. Haraldson T, Karlsson U, Carlsson GE. Bite force and oral function in complete denture wearers. J Oral Rehabil. 1979. 6:41–48.
41. Haraldson T, Carlsson GE, Ingervall B. Functional state, bite force and postural muscle activity in patients with osseointegrated oral implant bridges. Acta Odontol Scand. 1979. 37:195–206.
42. Mericske-Stern R, Zarb GA. In vivo measurements of some functional aspects with mandibular fixed prostheses supported by implants. Clin Oral Implants Res. 1996. 7:153–161.
TOOLS
Similar articles