0% found this document useful (0 votes)
80 views144 pages

Aditya LD

Uploaded by

dradibanik
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
80 views144 pages

Aditya LD

Uploaded by

dradibanik
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 144

LIBRARY DISSERTATION FOR M.D.S.

IN
PROSTHODONTICS AND CROWN AND BRIDGE

BIOMECHANICAL CONSIDERATIONS IN IMPLANT


DENTISTRY

SESSION: 2018-2021

DR. ADITYA BANIK


REGISTRATION NO.: 201800118

DEPARTMENT OF PROSTHODONTICS AND CROWN


AND BRIDGE

GURU NANAK INSTITUTE OF DENTAL SCIENCES AND


RESEARCH,
157/F, NILGUNJ ROAD, PANIHATI, KOLKATA-700114
LIBRARY DISSERTATION FOR M.D.S. IN
PROSTHODONTICS AND CROWN AND BRIDGE

BIOMECHANICAL CONSIDERATIONS IN IMPLANT


DENTISTRY

SESSION: 2018-2021

DR. ADITYA BANIK


REGISTRATION NO.: 201800118

DEPARTMENT OF PROSTHODONTICS AND CROWN


AND BRIDGE

GURU NANAK INSTITUTE OF DENTAL SCIENCES AND


RESEARCH,
157/F, NILGUNJ ROAD, PANIHATI, KOLKATA-700114
THE WEST BENGAL UNIVERSITY OF HEALTH SCIENCES

PROFORMA FOR SUBMITTING LIBRRARY DISSERTATION FOR MDS IN


DEPARTMENT OF PROSTHODONTICS AND CROWN AND BRIDGE

NAME OF THE STUDENT: ADITYA BANIK


NAME OF THE SUBJECT: BIOMECHANICAL CONSIDERATIONS IN IMPLANT
DENTISTRY

NAME OF THE GUIDE:


PROF. (DR.) JAYANTA BHATTACHARYYA (H.O.D. AND PRINCIPAL)
B.D.S. (CAL), M.D.S. (CAL)
DEPT. OF PROSTHODONTICS AND CROWN & BRIDGE,
GURU NANAK INSTITUTE OF DENTAL SCIENCES AND RESEARCH,
157/F, NILGUNJ ROAD, PANIHATI, KOLKATA – 700114.

NAME OF THE CO-GUIDE 1:


PROF. (DR.) SAMIRAN DAS
B.D.S. (N.T.R.U.H.S.), M.D.S. (A.I.I.M.S.)
DEPT. OF PROSTHODONTICS AND CROWN & BRIDGE,
GURU NANAK INSTITUTE OF DENTAL SCIENCES AND RESEARCH,
157/F, NILGUNJ ROAD, PANIHATI, KOLKATA – 700114.

NAME OF THE CO-GUIDE 2:


DR. SAYAN MAJUMDAR
READER,
B.D.S. (N.U.), M.D.S. (R.G.U.H.S.)
DEPT. OF PROSTHODONTICS AND CROWN & BRIDGE,
GURU NANAK INSTITUTE OF DENTAL SCIENCES AND RESEARCH,
157/F, NILGUNJ ROAD, PANIHATI, KOLKATA – 700114.
FORWARDED

SIGNATURE OF THE CANDIDATE:

APPROVED:
SIGNATURE OF THE GUIDE:
With official seal and date

SIGNATURE OF THE CO-GUIDES:


With official seal and date

1.

2.

COUNTERSIGNED:
SIGNATURE OF THE HEAD OF THE DEPARTMENT:
With official seal and date

COUNTERSIGNED:
SIGNATURE OF THE HEAD OF THE INSTITUTION
With official seal and date
BIOMECHANICAL CONSIDERATIONS IN IMPLANT DENTISTRY

LIBRARY DISSERTATION SUBMITTED

TO

THE WEST BENGAL UNIVERSITY OF HELATH SCIENCES

IN THE PARTIAL FULFILLMENT FOR THE AWARD OF THE DEGREE OF

M.D.S.

(MASTER OF DENTAL SURGERY)

IN

PROSTHODONTICS AND CROWN & BRIDGE

DR. ADITYA BANIK

REGISTRATION NO.: 201800118

THE WEST BENGAL UNIVERSITY OF HELATH SCIENCES

2018-2021
ACKNOWLEDGEMENTS

THANK YOU GOD

I would like to owe my first debt of gratitude to my guide, Prof. (Dr.) Jayanta Bhattacharyya,
Head of Department of Prosthodontics and Crown & Bridge & Principal, GNIDSR who has
guided me through with supervision, guidance and encouragement which have been very
valuable to me.

I am equally grateful to my co-guides Prof. (Dr.) Samiran Das and Dr. Sayan Majumdar who
have been a constant source of knowledge and encouragement and a complete driving force
since the very beginning.

I would also like to thank Prof. (Dr.) Soumitra Ghosh and (Dr.) Preeti Goel. Their contribution
has undoubtedly been a part of my successful work.

I would also like to thank Dr. Kritika Rajan for her invaluable assistance.

I would also like to acknowledge and thank my seniors, Dr. Ankita, Dr. Soumadip, Dr. Sthita,
Dr. Sudipto. Dr. Dhiman, Dr. Sreya, Dr. Partha, Dr. Anuradha, Dr. Saumyadeep, Dr. Supriyo.

I would also like to thank my colleagues Dr. Shraddha, Dr. Nandana, Dr. Sujoy & Dr. Sanjukta.

I would be failing in my duties if I don’t thank my wonderful friends Dr. Ritashna, Dr. Gauri,
Dr. Garimaa, Dr. Sushmit, Dr. Veerendar, Dr. Aishwarya, Dr. Sharanya, Dr. Shromi & Dr.
Pritam for their constant love, care and support and good wishes.

I would also take this opportunity to thank Mr. Agnitra Ganguly for his constant positivity,
direction & encouragement.

Lastly, I owe my most valued gratitude to my parents, Mrs. Promita Banik, Mr. S.N. Banik and
also to my elder sister Ms. Soumi Banik, without whose love, sacrifice and prayers, I would
not have been able to attain any possibility. I shall be forever grateful and indebted to them.

Aditya Banik

i
CONTENTS

I. INTRODUCTION……………………………………………………...1-8

II. TERMINOLOGIES AND CONCEPTS ASSOCIATED WITH DENTAL


IMPLANT BIOMECHANICS……...………………………………...9-16

III. REVIEW OF LITERATURE……………….……………………….17-33

IV. TOOTH VERSUS DENTAL IMPLANT SUPPORT SYSTEM…….34-40

V. COMPONENTS OF FORCE APPLIED TO DENTAL


IMPLANTS………………………………………………………….41-58

VI. BIOMECHANICAL RESPONSE AT DENTAL IMPLANT-TISSUE


INTERFACE………………………………………………………...59-77

VII. FORCE DELIVERY AND FAILURE MECHANISMS…………….78-86

VIII. OVERLOADING RISK FACTORS FOR DENTAL IMPLANTS….87-95

IX. SCIENTIFIC RATIONALE FOR IMPLANT DESIGN AND


PROSTHESIS SELECTION………………………………………96-122

X. TREATMENT PLANNING BASED ON BIOMECHANICAL


FACTORS………………………………………………………...123-124

XI. CONCLUSION…………………………………………………….….125

XII. ACKNOWLEDGEMENTS…………………………………. i

XIII. CONTENTS…………………………………………………. ii

XIV. LIST OF FIGURES………………………………………… iii-vii

XV. BIBLIOGRAPHY…………………………………………… viii-xiv

ii
LIST OF FIGURES

Chapter 1 – INTRODUCTION

Fig 1.1 – Completely edentulous maxillary and mandibular arches.

Fig 1.2 – Completely edentulous maxillary arch with partially edentulous mandibular arch.

Fig 1.3 – Deep dental caries (G.V. Black classification, Class -I) in respect to 36.

Fig 1.4 – Grossly decayed tooth, Root stumps in respect with 16.

Fig 1.5 – Severe Generalised Chronic Periodontitis.

Fig 1.6 – X-ray revealing the titanium optic chamber embedded in the tibia of a rabbit by Dr.
Branemark (1952).

Fig 1.7 – Schematic representation of an endo-osseous dental implant replacing a natural tooth.

Chapter 2 – TERMINOLOGIES AND CONCEPTS ASSOCIATED WITH DENTAL


IMPLANT BIOMECHANICS.

Fig 2.1 – The different types of force/stress.

Fig 2.2 – Stress strain relationship graph/ curve. Stress (Y- axis) and Strain (X- axis).

Fig 2.3 – Force couple/ pure moment.

Fig 2.4 – A diagrammatic representation of finite element analysis (FEM) of stress


concentration at the implant- abutment junction.

CHAPTER 4 – TOOTH VERSUS IMPLANT SUPPORT SYSTEM.

Fig 4.1 – A schematic representation of a natural tooth and a dental implant.

iii
Chapter 5 – COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS.

Fig 5.1 – The six different types of force (Apical and Occlusal, Lingual and Facial, Mesial and
Distal) a dental implant is subjected to in a three- dimensional plane i.e. Faciolingual,
Mesiodistal, Apicocclusal.

Fig 5.2 – A schematic diagram of a tooth loaded with 44.5 N force (f) acting along the point
A.

Fig 5.3 – Vector components of the applied force (f) acting on the point B, in the three planes.

Fig 5.4 – Multiple forces acting on single point (F1 F2 F3) with the resultant force (FR) is the
vectoral sum.

Fig 5.5 – Schematic representation of Moment/torque at three different situations.


(Left) The typical seesaw situation in which one person is 35kgs and another being 70kgs with
each being 2 metre and 1 metre away from the CR to have a balanced moment.
(Centre) Force (F)acting on an implant at the point C parallel to the long axis of the implant
with (d) being the distance in between them, thus Moment at point B, MB = F d
(Right) Similarly F acting perpendicular on the long axis of natural tooth, the Moment at the
Centre of Rotation CR, MCR= Fd
Fig 5.6 – Deflection and strain with FN being the perpendicular component of F and FS being
the shear component. ∆𝑙 being the deformation in length.

Fig 5.7 – Stress (Load, force) – Strain(deformation) relationship plotted against a graph with
stress (Y axis) and strain (X axis).

Fig 5.8 – Force (F) acting on the anterior region of the mandible with the Condyle (Centre of
Rotation CR) and M1 (lateral pterygoid) and M2 (Masseter) are the two important balancing
muscles.

Fig 5.9 – Rangert Model with two implants #1 and #2 with distance between them being b, and
the distance between #2 and the terminal end of the bridge a.

Fig 5.10 - Rangert Model with four implants #1 #2 #3 #4.

Fig 5.11 and 5.12 – Skalank Model with six implants (1, 2, 3, 4, 5,6) and four implants, two
removed from the previous scenario model (2, 3, 4, 5).

Fig 5.13 – Rangert Model with two implants #1 and #2, with #1 being tilted at a known
angulation (30O).

iv
Fig 5.14 –Skalank Model showing rigid parameters as considerations for analysis.

Fig 5.15 – FR acting as a vector at different directions on structure composed of two different
component materials. FN being the perpendicular component and FS being the shear
component.

Chapter 6 – BIOMECHANICAL RESPONSE AT DENTAL-IMPLANT TISSUE


INTERFACE

Fig 6.1 – A schematic representation of the various stages/steps in bone remodelling.

Fig 6.2 – (Left) Interfacial fibrous membrane of a rough-surfaced hydroxyapatite implant that
underwent 150 microns of micromotion in a dog femur during a healing period of 4 weeks.
(Right) Direct bone- hydroxyapatite interface under the absence of micromotion (Soballe et al
1992)

Fig 6.3 – Scanning Electron Microscopy (SEM) of bone/ DAE treated Ti6Al4V implant
interface. (Top left) showing low magnification showing bone-implant interface. (Top right)
bone implant interface showing cement line as globular accretions deposited on the surface of
the implant. (Bottom left) different sizes of globular accretions deposited within the complex
topography of the implant surface. (Bottom right) high magnification of the globular
accretions.

Fig 6.4 – A schematic representation of Wolff’s law applied to femur with the load being the
weight of the body.

Fig 6.5 – A study showing a finite element analysis (FEA) of stress distribution at the dental
implant-bone interface with the implant being placed at different angulations and height.

Fig 6.6 – Implant-abutment/prosthesis interface.

Chapter 7 – FORCE DELIVERY AND FAILURE MACHANISM

Fig 7.1 – A force couple acting on a dental implant supported cantilever bridge with two
moment arms at the implant and at the terminal pontic. The distance between the two moment
arms (1.5cm) increases the moment load to 150 N-cm from 100 N.

v
Fig 7.2 – Various force/moment loads acting on a dental implant at the three different planes
and their various manifested movements.

Fig 7.3 – Two of the three clinical moment arms, i.e. (Top) Cantilever Length. (Bottom)
Occlusal height.

Fig 7.4 - Clinical moment arm cantilever length.

Fig 7.5 - (A) Graph plotted for applied stress (Y- axis) versus number of loading cycles (Y-
axis). (B) Fatigue behaviour of two biomaterials i.e. Commercially Produced Titanium versus
Ti-6Al-4v.

CHAPTER 8 – OVERLOADING RISK FACTORS FOR DENTAL IMPLANTS.

Fig 8.1 – Bruxism habit causing generalised attrition.

Fig 8.2 – Complex tongue thrusting.

Fig 8.3 – A schematic representation of an implant supported overdenture opposing natural


dentition.

Fig 8.4 – A Finite Element Analysis of occlusal load acting on dental implant with different
crown height.

Fig 8.5 – Lekholm and Zarb (1985) classification of different type of quality and quantity of
bone.

Fig 8.6 – Implant supported cantilever bridge.

Fig 8.7 – Pier abutment (the premolar).

Fig 8.8 – Bending/ lateral displacement of an implant due to masticatory forces.

Chapter 9 – SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS


SELECTION.

Fig 9.1 - A smooth surfaced implant.

Fig 9.2 – Different implant body designs.

Fig 9.3 – A diagrammatic representation of a threaded implant.

vi
Fig 9.4 – Different forces acting on two different body designs of a possible dental implants
(V-shaped and Square shaped).

Fig 9.5 – The different shapes/designs of a crest module in an implant. (Right) straight crest
module (Centre) converging angled crest module (Left) diverging angled crest module.

Fig 9.6 – Two different implants with different pitch, (Right) Coarser pitch (Left) Finer pitch.

Fig 9.7 – A schematic representation of a gold screw tightening leading to a disbalance or


bending of the final prosthesis.

Chapter 10 – TREATMENT PLANNING BASED ON BIOMECHANICAL FACTORS

Fig 10.1 – (Right) Implants placed offset to the centre of the prosthesis in a tripod arrangement.
(Left) Excessive height of restoration causing unwanted axes of force.

Fig 10.2 – Different cuspal inclination of a prosthesis with the direction of the force acting on
it.

vii
INTRODUCTION

An adequate dentition is of importance for well-being and life quality. Despite advances
in preventive dentistry, edentulism is still a major public health problem worldwide.
Edentulism is a debilitating and irreversible condition and is described as the ‘final marker of
disease burden for oral health’. Although the prevalence of tooth loss has declined over the last
decade, complete or partial edentulism remains a major disease worldwide, especially among
older adults.

Fig: 1.1

Fig : 1.2

Dental caries and periodontal disease have historically been considered the most
common causes of complete or partial edentulism. The other common causes for tooth loss
include poor oral hygiene, trauma, sports injury, bruxism, jaw surgery, traumatic occlusion,

INTRODUCTION 1
eating disorders, root perforation, genetic predisposition, congenital defect, systemic disease
(like diabetes) and lack of nutrients.

Fig : 1.3 Fig : 1.4

Fig : 1.5

The consequences associated with being partial/complete edentulous are listed below:

• Loss of function.
• Loss of aesthetics.
• Loss of facial support and masticatory insufficiency.
• Problems with pronunciation and phonetics.
• Eating insufficiency.
• Impede normal contour and comfort.
• There may be drifting and tilting of adjacent teeth.
• Supra-eruption of opposing teeth.
• Temporomandibular joint disorders.

INTRODUCTION 2
Monitoring this occurrence of an oral ‘end state’ of edentulism is important because it
is an indicator of both population health and the functioning and adequacy of a country's oral
health care system.

Wu B et al (2012) found a wide variation in edentulism prevalence among adults aged


50 and above in five ethnic groups: Asians, African Americans, Hispanics, Native Americans,
and non-Hispanic Caucasians. In 2008, Native Americans had the highest predicated rate of
edentulism based on oral exam (24%), followed by African Americans (19%), Caucasians
(17%), Asians (14%), and Hispanics (14%). Another study in 2009, UK Adult Dental Health
Survey self-reported that the percentage of adults who were edentulous was 5% for those aged
55–64 years and 15% for those aged 65–74 years. Another study by Peltzer K et al (2014),
prevalence of edentulism was investigated in six countries China, Mexico, Russian Federation,
South Africa, India, Ghana and it was found that countries like India, Mexico and Russia than
China, Ghana and South Africa. This study was basically to evaluate and identify factors
associated with the presence of edentulism among older adults across the six countries help to
identify areas for further exploration and targets for intervention.

Although the prevalence and patterns of tooth loss have been extensively studied in the
western countries, very few such studies have been conducted in India. An institutional based
survey conducted by Sonkesariya S et al (2014) revealed that nearly 38.5 and 63.1% and 61.4
and 36% of urban and rural female and male population was dentulous in the age of gap 53 and
43 years respectively. Another study conducted by Jandial S et al (2017) revealed complete
edentulism in the following age group:

AGE GROUP MALE FEMALE


45-55 years old. 55.1 % 58.8%
55-65 years old. 54.7% 32.6%
Above age of 65 years old 65.2% 64.5%

It was also seen in partially edentulous patients in the following age group:

AGE GROUP FEMALE MALE


45-55 years old. 32.6 % 33.3%
55-65 years old. 63.5% 58.5%
Above age of 65 years old 35.4% 34.8%

INTRODUCTION 3
In another study by Madhankumar S et al (2015), the prevalence of partially
edentulous condition was found to be:

KENNEDY’S FEMALE MALE


CLASSIFICATION
CLASS I 11% 9%
CLASS II 10% 14%
CLASS III 49% 51.4%
CLASS IV 13% 3%

It can be concluded that there is highest prevalence of Kennedy’s Class III partially
edentulous situation in both males and females, and the least in Kennedy’s Class I in females
and Kennedy’s Class IV in males.

From the findings of these studies, it can be concluded that the prevalence of
edentulousness increases with age which results in various long-term effects of tooth removal
on patient’s facial structure and general well-being, thereby increasing the need for
prosthodontics rehabilitation. Awareness and proper dental education regarding proper dental
hygiene and timely replacement of the missing teeth need to be taken care of.

Edentulism exists, it will remain prevalent, and its management is beneficial to the
affected population and society. Although the most traditional method of rehabilitation in
edentulous patients is complete dentures, it is further associated with an array of related
complications and associated clinical manifestations of denture use as following:

• Denture Stomatitis.
• Traumatic Ulcers.
• Irritation-Induced Hyperplasia.
• Altered Taste Perception.
• Burning Mouth Syndrome.
• Gagging.
• Patient Discomfort.

INTRODUCTION 4
A complete revolution was brought into the field on dentistry with the advent of
implants. Dr Per-Ingvar Branemark (1952), a Swedish physician surgically placed titanium
optic chamber in the tibia of a rabbit and studied it at a microscopic level. He observed that the
titanium chambers were inseparably incorporated within the bone tissues, which actually grew
into very thin spaces in titanium. Although the phenomenon of osseointegration was first
described by Bothe et al. (1940) and later by Leventhal et al. (1951) it was later popularised
by Branemark and it was found that a foreign implant can be used for bone anchorage due to
the phenomenon of ‘biological fusion’ or osseointegration.

OSSEOINTEGRATION :
1. the apparent direct attachment or connection of osseous tissue to an inert, alloplastic material
without intervening fibrous connective tissue;
2. the process and resultant apparent direct connection of an exogenous material’s surface and
the host bone tissues, without intervening fibrous connective tissue present;
3. the interface between alloplastic materials and bone. (GPT: 9).

IMPLANT : To graft or insert a material such as an alloplastic substance, an encapsulated drug,


or tissue into the body of a recipient. (GPT: 9)

IMPLANT : Any object or material, such as an alloplastic substance or other tissue, which is
partially or completely inserted or grafted into the body for therapeutic, diagnostic, prosthetic,
or experimental purposes. SYNONYM - DENTAL IMPLANT. (GPT: 9)

Fig : 1.6

INTRODUCTION 5
The growth of osseointegrated implants symbolizes one of the most significant
breakthroughs in current dental practice and in the oral rehabilitation of partially or completely
edentulous patients. A shift towards improved aesthetics and simplified use has resulted in the
application of oral implants in the replacement of single teeth from the conventional prosthesis.

The advantages of prosthodontic therapy using oral implants from the conventional
methods of removable partial denture and complete denture prosthesis are as follows: -

• Increased stability.
• High Patient compliance.
• More comfort.
• Increased masticatory function
• Induction of bone apposition and prevents bone resorption.
• More aesthetic.
• Avoidance of food impaction
• Better oral hygiene maintenance.
• Improved speech and phonetics.
• Longevity is high.

Fig : 1.7

INTRODUCTION 6
The word ‘BIOMECHANICS’ is delivered from the ancient Greek word bios which
means ‘life’ and mechanike which means ‘mechanics’, thus it refers to the study of the
mechanical principles of living organisms, particularly their movement and structure.

BIOMECHANICS :

1. the application of mechanical laws to living structures, specifically the locomotor systems
of the body;

2. the study of biology from the functional viewpoint;

3. an application of the principles of engineering design as implemented in living organisms;


SYNONYM - DENTAL BIOMECHANICS (GPT: 9)

The discipline of biomedical engineering, which applies engineering principles to


living systems, has unfolded a new era in diagnosis, treatment planning, and rehabilitation in
patient care. One aspect of this field, biomechanics, concerns the response of biological tissues
to applied loads. Biomechanics uses the tools and methods of applied engineering mechanics
to search for structure–function relationships in living materials. It comprises the response of
the biologic tissues to the applied loads.

Usually, in any structure subjected to functional loads, there may be situations leading
stress and strain and eventually if not controlled may cause overload and subsequent
complications. When we come to the biomechanics of dental implants, here the implant
treatment defines a structure based on both the biologic tissues (bone) and the mechanical
components (implant and superstructure). So, the biomechanics of dental implants concerns
the response of biologic tissues to the applied forces and loads. Hence for the better planning
and ultimate success of dental implants it is very necessary to understand the basic of
biomechanics and its importance.
The absence of a periodontal ligament around the dental implant reduces proprioception
and the patient's reflex function, as well as the implant not being able to migrate to compensate
for premature occlusal contacts like natural dentition with a periodontal ligament. Implants and
their rigid-attached restoration need to render any movements like that of the natural teeth.
Therefore, the biomechanical assessment is very crucial for implant success.

INTRODUCTION 7
The importance of biomechanics in dental implants was initially not studied
extensively. Clinical experience and research over the years has shown the significant
importance of biomechanics in the success and predictability of implants. Thus, this library
dissertation intends to provide a better understanding of biomechanics and presents an
overview about its role and considerations in implant dentistry.

INTRODUCTION 8
TERMINOLOGIES AND CONCEPTS ASSOCIATED
WITH IMPLANT BIOMECHANICS

BIO-MECHANICS :

Application of mechanical principles and design to biologic structure; the interface


between biology and mechanics.

FORCE :

It is defined as an agency that, when exerted on a body tends to set the body into motion
or to alter its present state of motion.

Force applied to any material causing deformation of that material.

Units : Newton (N). S.I.Unit

Formula for calculating force is f = ma, where f = force, m = mass, a = acceleration due
to gravity (9.8m/s²). Being vector force has definite magnitude, a specific direction and a point
of application. Forces may be described by the magnitude, duration, and direction type and
magnification factors.

Forces acting on dental implants are vector quantities. (magnitude and direction).
Maximum bite forces exhibited by adults can be affected by the following:

1. Age.
2. Sex.
3. Degree of edentulousness.
4. Bite location.
5. Parafunctional habits.

TYPES OF FORCES :

Compressive Forces: The internal induced force that opposes the shortening of a
material in a direction parallel to the direction of the stresses; any induced force per unit area

TERMINOLOGIES AND CONCEPTS 9


that resists deformation caused by a load that tends to compress or shorten a body. These forces
tend to maintain the integrity of implant interface.

Tensile forces: The internal induced force that resists the elongation of a material in a
direction parallel to the direction of the stresses.

Shear forces: The internal induced force that opposes the sliding of one plane on an
adjacent plane or the force that resists a twisting action.

STRESS :

It is defined as force per unit area. It is normally defined in terms of mechanical stress.
The manner in which a force is distributed over a surface is referred as mechanical stress.
Treatment planning should be made to minimize and evenly distribute the mechanical stress
on the implant structure and underlying continuous bone. The magnitude of stress depends on
the following:

1.Force.

2.Cross-sectional area.

The magnitude of the force can be reduced by controlling the following factors:

1. Cantilever length.
2. Offset loads.
3. Crown height.
4. Night guards for patients with Para-functional habits.
5. Over dentures rather than fixed prosthesis.

Functional cross sectional area:

• It is defined as that surface that participates in load bearing and stress dissipation.

σ = f/a,

σ = stress (psi),

f = force (N),

a = area (sq. inches).

TERMINOLOGIES AND CONCEPTS 10


TYPES OF STRESSES ASSOCIATED WITH IMPLANT BIOMECHANICS ARE :

Compressive Stress: It is defined as the internal induced force that opposes the shortening of
a material in a direction of the stresses, any induced force per unit area that resists deformation
caused by a load that tends to compress or shorten a body.

Shearing Stress: It is the internal induced force that opposes the sliding of one plane on to an
adjacent plane or the force that resists twisting action.

Tensile Stress: It is the internal induced force that resists the elongation of a material in a
direction parallel to the direction of the stress.

Fig: 2.1

STRAIN: It is defined as change in length per unit length when stress is applied.

TERMINOLOGIES AND CONCEPTS 11


STRESS AND STRAIN RELATIONSHIP

The stress and strain curve is constructed by plotting stress areas along the vertical axis and
strain along the horizontal axis.

Fig: 2.2

Typical stress-strain curve illustrating parameters used for comparison of mechanical


responses of materials.

PL - PROPORTIONALITY LIMIT,

EL - ELASTIC LIMIT,

Q ys - YIELD STRENGTH,

Q max - MAXIMUM STRESS,

E max - MAXIMUM STRAIN.

MODULUS OF ELASTICITY :

It is the ratio of stress to strain in the linear portion of the stress-strain curve and is a
measure of stiffness of that material.

TERMINOLOGIES AND CONCEPTS 12


PROPORTIONAL LIMIT :

When a material is subjected to increasing stress, a point is reached where the stress is
no longer proportional to strain, this stress is called proportional limit.

ELASTIC LIMIT :

It is defined as the stress beyond which permanent deformation results.

YIELD STRENGTH :

It is defined as the stress corresponding to a specific amount of permanent deformation.


Permanent deformation is usually taken to be 0.1% or 0.2% strain and is referred to as percent
offset. A material with high yield strength is more difficult to permanently deform than a
material with low yield strength.

ULTIMATE STRENGTH :

With continued loading the specimen will fracture. The stress in the specimen at the
instant of fracture is called the ultimate strength.

MAXIMUM ELONGATION :

It is a measure of how much change in shape the specimen underwent at the time of
fracture. It is commonly expressed in terms of percentage strain.

COUPLE :

Couple is a pair of concentrated forces having equal magnitude and opposite direction
with parallel but non-collinear line of action. A couple when acting on a body brings about
pure rotation.

TERMINOLOGIES AND CONCEPTS 13


Fig : 2.3

MOMENT :

Moment can be defined as the measure of rotational potential of a force with respect to
a specific axis. Units of measurement of moment are -gram millimeters.

HARDNESS :

• It is the most common parameter used to compare dental materials.

• It is defined as the resistance of material to fracture or to an indentation or to a scratch.

CREEP :

It is denoted by time dependent plastic deformation.

RELAXATION :

It is the decay of stresses within a material when subjected to constant deformation.

TERMINOLOGIES AND CONCEPTS 14


METHODS USED TO STUDY STRESS ANALYSIS

The techniques of stress analysis can be separated into theoretical and experimental
subgroups. The theoretical techniques use mathematical formulations and solutions of the
resulting equations. The experimental techniques usually involve measurements of various
types made directly on the structures of interest or through the use of modelling of the structure.

THEORETICAL TECHNIQUES :

Theoretical techniques involve using the basic laws of physics (e.g. conservation of
momentum, conservation of energy, conservation of mass) and the equations that specify the
stress-strain relationship of the materials which is made to formulate the governing differential
or integral equations for the structure. These equations are then solved by using numeric or
analytical methods.

Finite Element Analysis :

This technique involves visualization of structure into one by assembling a finite


number of discrete structural elements connected at a finite number of points. It can also be
defined as a numerical method of solving systems of partial differential equations (PDEs). It
reduces a PDE system to a system of algebraic equations that can be solved using traditional
linear algebra techniques. In simple terms, FEM is a method for dividing up a very complicated
problem into small elements that can be solved in relation to each other.

A finite element model is constructed by dividing solid objects into several elements
that are connected at a common nodal point. Each element is assigned appropriate material
properties corresponding to the properties of the object being modelled. The first step is to
subdivide the complex object geometry into a suitable set of smaller ‘elements’ of ‘finite’
dimensions. When combined with the ‘mesh’ model of the investigated structures each element
can adopt a specific geometric shape (i.e., triangle, square, tetrahedron, etc.) with a specific
internal strain function. Using these functions and the actual geometry of the element, the
equilibrium equations between the external forces acting on the element and the displacement
occurring at each node can be determined.

TERMINOLOGIES AND CONCEPTS 15


This is the most commonly used analysis to study stress, strain and their relationship
in respect to implant biomechanics although other mentionable analysis are photoelastic and
strain gauge.

Finite element analysis is a numerical stress analysis technique and is extensively


used in implant dentistry to evaluate the risk factors from a biomechanical point of view.
Simplifications and assumptions are the limitations of finite element analysis (FEA)studies.
Although advanced computer technology is used to obtain results from simulated models,
many factors affecting clinical features such as implant macro and micro design, material
properties, loading conditions, and boundary conditions are neglected or ignored. Therefore,
correlating FEA results with preclinical and long-term clinical studies may help to validate
research models.

Fig : 2.4

TERMINOLOGIES AND CONCEPTS 16


REVIEW OF LITERATURE

Borchers L, Reichart P (1983) : In their study, subjected an anchor-type ceramic dental


implant contained in a section of the mandibular molar region for finite element stress analysis
(FEM). The distributions of stresses in the bone around the implant due to axial and transverse
loading were calculated for different stages of normal and pathological development of the
implant-bone interface. From the results of FEM stress analysis of an anchor type A1203-
ceramic dental implant, the following conclusions may be drawn:

1. High stress peaks which were calculated in the crestal region of the alveolar bone, especially
with transverse loading, might cause bone resorption, connective tissue ingrowth, and
subsequent implant failure. Stress concentrations were most distinct with the implant's
surroundings consisting of cancellous bone.

2. Presence in the model of a lamina dura or a connective tissue layer around the implant was
found to reduce stress peaks.

3. Further analysis should be done to evaluate design improvements in order to increase the
share of load transmitted by the implant wings, thus unburdening the stem region.

Albertson T, Zarb G, Worthington (1986) : They reported that factors such as surgical
technique, host bed, implant design, implant surface, material biocompatibility, and loading
conditions have been shown to affect implant osseointegration. They also stated that
characteristics of macro-design may improve the primary stability of dental implants and
decrease the stress concentration on the implant body and the superstructure.

Kinni M E et al (1987) : They studied the the stress distributing characteristics of the Core-
Vent and Brånemark implants, that were done using the techniques of quasi three-dimensional
photoelasticity. With the assumption of complete osseointegration, it was shown that the
Brånemark implant distributes axial and inclined loads to the supporting structures in a more
equitable and potentially more biologically acceptable manner than Core-Vent implants.

REVIEW OF LITERATURE 17
Brunski J B (1988) : He studied and gave a review about an insight into the design process
and biomaterial/biomechanical aspects of endosseous implant design and specific facets that
are to be considered related to materials, implant shape, special surface coatings, shock-
absorbers, and the implant-tissue interface.

Setz J et al (1989) : They conducted a study to measure the implant stress during a chewing
cycle for implant supported complete dentures using ‘electrognathographic’ measurements on
18 patients. Patients with implant-attached mandibular complete dentures showed larger and
more stable chewing patterns compared to their former dentures without implants. Strain gauge
measurements revealed mechanical stress acting on the implants even during the fixation of
the bar. In addition to simulating the periodontal ligament, the intramobile element (IME) of
the IMZ system compensated minor, clinically undetectable mismatches inherent with
technical procedures (casting, soldering). Although the implants are loaded with tensile and
compressive stress, tensile stress dominated during the chewing process. The implants were
also loaded during swallowing. This load reached half of the value of the chewing loads. It was
concluded that the retention of a mandibular complete denture by two intermorainal inserted
IMZ implants and a clip-bar attachment provides the patient with a more secure feeling. This
was proven by chewing patterns (Sirognathograph recordings) of 18 equivalently treated
patients. With the dentures attached to the bar, chewing movements became wider and centric
relation was reached earlier as compared to the same test without a bar. Both facts indicate a
more effective masticatory function.

Cook S D et al (1989) : They conducted a study using a three-dimensional finite element


analysis (FEA) to determine the effect of implant elastic modulus of stresses in tissues around
LTI carbon and aluminium oxide dental implants. The finite element model was constructed
using a baboon mandible containing a blade type dental implant. A three-unit fixed bridge was
modelled connecting the dental implant to a natural molar. The results of the study indicate that
stress levels of approximately a factor of 3 lower in the crestal region can be expected for
aluminium oxide implants when compared to the LTI carbon implants. It was also observed
that the use of LTI carbon and aluminium oxide dental implants as an abutment in a fixed
bridge results in a reduction of stresses in tissues around the natural tooth when compared to
normal physiological stress levels.

REVIEW OF LITERATURE 18
Van Rossen I P et al (1990) : They studied with the means of finite element analysis,
calculations were made of the stress-distribution in bone around implants with and without
stress-absorbing elements. A free standing implant and an implant connected with a natural
tooth were simulated. For the freestanding implant, it was concluded that variation in the elastic
modulus of the stress-absorbing element had no effect on the stresses in bone. Changing the
shape of the stress-absorbing element had little effect on the stresses in cortical bone. For the
implant connected with a natural tooth, it was concluded that a more uniform stress was
obtained around the implant with a low elastic modulus of the stress-absorbing element. It was
also concluded that the bone surrounding the natural tooth showed a decrease in the height of
the peak stresses.

Gallas M M et al (1991) : The aim of their study was to examine the bone and the implant
finite element models. The first model was considered that there was no osseiointegration and
the second model to be with complete osseointegration. The models were used to study the
distribution of stress transfer and its pattern with ITIR endosseous implants and its supporting
tissues. Their study also examined threaded implant placed in an edentulous segment of a
human mandible with cortical and cancellous bone. The results concluded was that both models
indicated maximum stress was always located around the neck of the implant, in the marginal
bone. Thus, this area should be preserved clinically in order to maintain the bone implant
interface structurally and functionally.

Meijer G et al (1992): They studied the stress distribution around dental implants was by using
of a two-dimensional model of the mandible with two implants. A vertical load of 100 N was
imposed on abutments or the bar connection. The stress was calculated for a number of
superstructures under different loading conditions with the help of the finite element method.
The length of the implants and the height of the mandible were also varied. It was seen a model
with solitary abutments showed a more uniform distribution of the stress when compared with
a model with connected abutments. The largest compressive stress was also less in the model
without the bar. Using shorter implants did not have a large influence on the stress around the
implants. When the height of the mandible was reduced, a substantially larger stress was found
in the bone around the implants because of a larger overall deformation of the lower jaw.

REVIEW OF LITERATURE 19
Clift S E et al (1992): In this study, a dental implant which had the same geometry as the
Branemark system, but with a bioactive surface coating added to produce a direct bond to the
bone, was analysed. A finite element stress and strain analysis has been carried out for a range
of bone density distributions under axial and lateral loading. The predictions indicated that
there was no evidence of strain shielding around the neck of the implant. With lateral loading,
high values of von Mises stresses (18 MPa) were predicted around the neck of the implant. A
reduction in the elastic modulus of the bone around the neck of the implant by a factor of 16
only produced a two-fold reduction in the peak stress. This resulted in stress levels capable of
inducing fatigue failure in this much weaker bone. This analysis has demonstrated that it is
extremely important to have good quality dense bone around the neck of the implant to
withstand the predicted peak stresses of between 9 and 18 MPa. Failure to achieve this after
implantation and subsequent healing may result in local fatigue failure and resorption at the
neck upon resumption of physiological loading.

Lawrence A et al (1993): They reviewed and studied about and stated that the force distribution
between members of a system depends on a complex relationship between the relative stiffness
of the structural parts with its investment medium (periodontal ligament or osseointegration).
A rigid prosthesis is necessary to distribute force in all types of multiple-unit-supported
prostheses. When force is applied to one portion of a multiple-tooth-supported prosthesis, the
micromovement of the periodontal ligament (0.5 mm range) initiates movement of the whole
rigid structural entity (teeth and prosthesis). This micromovement distributes force to the
remaining natural teeth. With a multiple-implant-supported prosthesis, force application to one
portion is distributed to the nearest osseointegrated fixture interface. The force is concentrated
at that interface. The amount of distribution to the remaining fixtures depends on the degree of
deformation (flexibility) of the investing bone, fixture, abutment, retaining screws, and
prosthesis. Combined prostheses using implants and natural teeth should be approached with
caution. Internal attachments and/or telescopic coping construction have been used. However,
force transmission is completely different in both segments. Implants always support the
natural teeth, rather than visa-versa, because of the overwhelming differential in mobility
between periodontal ligament micromovement and the osseointegrated implant interface.

REVIEW OF LITERATURE 20
Rodriguez A et al (1994): In their review they summarised about the determination of an
acceptable length of a cantilever for fixed implant prosthesis. They discussed the effects of
biomechanical stress on the fixed implant prosthesis and supporting bone are central to the
development or implant prosthesis design.

Meijer G J et al (1995): They studied the influence of a three‐layered flexible coating of


Polyactive® on bone stress distribution was investigated by three‐dimensional finite element
models of mandibular bone, in which a titanium implant (coated or uncoated) was placed. Poly‐
active® is a system of poly (ethylene oxide) poly (butylene terephthalate) segmented co‐
polymers with bone‐bonding capacity. In the case of sagittal and transversal loading, the use
of a Polyactive® coating reduced both the minimum principal stress in the bone and the
compressive radial stress at the bone‐implant interface. However, it raised the maximum
principal and the tensile radial stress. In the case of vertical loading, the application of a flexible
coating reduced the compressive radial stress at the bone‐implant interface around the neck of
the implant by a factor of 6.6 and the tensile radial stress by a factor of 3.6 . Variations in
composition and thickness of the coating did not affect the results significantly.

Haack JE et al (1995) : They developed a new method to determine initial preload on UCLA-
type abutment screws by measuring elongation after applying known tightening torques with a
digital torque gauge. Loosening torque was also measured after tightening to 32 N-cm torque
for gold alloy abutment screws and 20 N-cm for titanium abutment screws. Gold alloy and
titanium abutment screws were each used to secure a gold UCLA hexed abutment to a titanium
implant. Stresses and forces were calculated from the elongation measurements for three
regions of each screw. Elongation of the screws after applying the manufacturer's
recommended tightening torques were within the elastic range. Induced stresses were
calculated to be at the value of 57.5 % and 56 % of the yield strengths for gold alloy and
titanium, respectively. It was concluded that tightening of screws beyond recommended levels
was possible without producing plastic deformation.

REVIEW OF LITERATURE 21
Luigi B et al (1996) : They conducted studies on the influence of implant diameter and its
length and its effect on stress distribution of osseointegrated implants related to the crestal
bone geometry using a three dimensional finite element analysis. It was found that maximum
stress areas were numerically located at the implant neck, and possible overloading could occur
in compression in compact bone (due to lateral components of the occlusal load) and in tension
at the interface between cortical and trabecular bone (due to vertical intrusive loading
components). Stress values and concentration areas decreased for cortical bone when implant
diameter increased, whereas more effective stress distributions for cancellous bone were
experienced with increasing implant length. For implants with comparable diameter and length,
compressive stress values at cortical bone were reduced when low crestal bone loss was
considered. Finally, dissimilar stress-based performances were exhibited for mandibular and
maxillary placements, resulting in higher compressive stress in maxillary situations. Thus it
was concluded that Implant designs, crestal bone geometry, and site of placement affect load
transmission mechanisms. Due to the low crestal bone resorption documented by clinical
evidence, the ankylossed implant based on the platform switching concept and sub-crestal
positioning demonstrated better stress-based performance and lower risk of bone overload than
the other implant systems evaluated.

Benzing U R et al (1997) : They studied the biomechanical aspects of these implant-prosthetic


concepts with clinical strain-gauge measurements and a theoretical three-dimensional analysis
was done using the finite element method. The results revealed that the distribution of bone
stresses is more favourable with a spread-out implant arrangement than with a concentrated
implant arrangement and cantilever restoration. The resistance to bending of a superstructure
has an influence on bone stress concentration that should not be ignored. Stresses are controlled
not only by the number or distribution of implants, but also by the material and design of the
superstructure.

Lai H et al (1998) : The objective was to study the stress around implants that may lead to
bone resorption and loss of the implant. The present study examined the influence of percentage
of osseointegration at the implant-bone interface on the transmission of occlusal forces for
endo-osseous dental implants. A three-dimensional finite element method used in the study
was built from data obtained from slices of dental computed tomography scans. The study

REVIEW OF LITERATURE 22
modelled a 3.75 x 10-mm cylindric implant placed in an edentulous mandible. Varying the
elastic parameters assigned to the implant-bone interface, a load of 35 N was applied at the
occlusal surface of the restoration at the vertical axis of the implant. Maximum principal stress,
minimum principal stress and Von Mises stress were calculated, and it was concluded that the
most extreme stresses in the bone were always located around the neck of the implant. Those
stresses in the implant-tissue interface decreased in inverse proportion to the increase in
percentage of osseointegration. These results indicate the value of osseointegration in the
aspect of mechanics.

Kim WD et al (1999) : They performed a study in which a photoelastic and strain gauge analysis
was performed to evaluate the stress transferred to implants through the provisional-cement-
retained, the permanent-cement-retained, and the screw-retained prostheses. The deflections of
the prostheses at the time of the loading were also measured. It was seen that in the single
crown test, the provisional-cement-retained crowns transferred less stress. In the two-unit fixed
partial denture test, there were no differences between the three different prostheses. In the
two-implant supported distal cantilevered prostheses, the screw-type and the permanent-
cement-retained prostheses developed more stress around the apex of both implants. The
permanent-cement-retained prostheses acted almost the same as the screw-type.

O’Mahony A, Bowles Q et al (2000) : The objective of this study was to evaluate the simulated
effects of axial and off-axial vertical loads on stress gradients at the implant/bone interface of
a single-unit osseointegrated root-form endo-osseous dental implant. A two-dimensional finite
element model was generated. A 490-N load was applied at 0, 2, 4, and 6 mm from the vertical
axis of the implant. Off-axis loading resulted in greatly increased compressive stresses within
the crestal cortical bone on the side to which the load was applied and similarly increased
tensile stresses on the side opposite the load. These stresses increased considerably with each
mm increase off axis of the applied load. It was concluded that off-axis loading of single-unit
implant restorations provides a significant contribution to increased stresses at the
implant/cortical bone interface. The distance off axis at which the load is applied is also
significant.

REVIEW OF LITERATURE 23
Himmlova L et al (2000): The study was conducted to understand the mathematical simulation
of stress distribution around implants and to determine which length and diameter of implants
would be best to dissipate stress. It was done using computations of stress arising in the implant
bed with finite element analysis, using 3-dimensional computer models. The models simulated
implants placed in vertical positions in the molar region of the mandible. A model simulating
an implant with a diameter of 3.6 mm and lengths of 8 mm, 10 mm, 12 mm, 14 mm, 16 mm,
17 mm, and 18 mm was developed to investigate the influence of the length factor. The
influence of different diameters was modelled using implants with a length of 12 mm and
diameters of 2.9 mm, 3.6 mm, 4.2 mm, 5.0 mm, 5.5 mm, 6.0 mm, and 6.5 mm. The masticatory
load was simulated using an average masticatory force in a natural direction, oblique to
the occlusal plane. Values of von Mises equivalent stress at the implant-bone interface were
computed using the finite element analysis for all variations. Values for the 3 most stressed
elements of each variation were averaged and expressed in percent of values computed for
reference (100%), which was the stress magnitude for the implant with a length of 12 mm and
diameter of 3.6 mm. The results concluded that the maximum stress areas were located around
the implant neck. The decrease in stress was the greatest (31.5%) for implants with a diameter
ranging from of 3.6 mm to 4.2 mm. Further stress reduction for the 5.0-mm implant was only
16.4%. An increase in the implant length also led to a decrease in the maximum von Mises
equivalent stress values; the influence of implant length, however, was not as pronounced as
that of implant diameter. At the end it was concluded that Within the limitations of this study,
an increase in the implant diameter decreased the maximum von Mises equivalent stress around
the implant neck more than an increase in the implant length, as a result of a more favourable
distribution of the simulated masticatory forces applied in this study.

Ciftci Y et al (2000) : They studied the effect of various materials used in fabricating
superstructures for implant retained fixed partial dentures on stress distribution around implant
tissues was investigated. Five different mathematical models consisting of 11,361 nodes and
54,598 elements were constructed to study porcelain, gold alloy, composite resin, reinforced
composite resin, and acrylic resin veneering materials using the 3-dimensional finite element
analysis method. MARC K7.2 Mentat 3.2 software was used for the analysis. Reference points
were determined on the cortical bone, where perpendicular, oblique, and horizontal forces were
applied. It was seen that the Stress values created by oblique and horizontal forces appeared to
be higher than those created by vertical forces. Stress seemed to be concentrated at the cortical

REVIEW OF LITERATURE 24
bone around the cervical region of the implant. Gold alloy and porcelain produced the highest
stress values in this region. Stresses created by acrylic resin and reinforced composite resin
were 25% and 15% less, respectively, than porcelain or gold alloy. Porcelain and gold alloy
produced stress values at the lingual implant sites that reached the ultimate strength values of
the cortical bone.

Anil N et al (2000) : The study investigated stresses formed around the implant and the
antagonist natural tooth under occlusal force in the substitution of a missing lower first molar
with a rigid or resilient IMZ (Intra Mobil Zylinder) using the FEM analysis method. The results
indicate that a bite force of 143 N resulted in high compressive stresses around the roots of a
natural tooth opposing a restoration supported by an IMZ implant with rigid abutment. It is
speculated that these high compressive stresses may contribute to intrusion of the tooth.

Cruz M et al (2001) : They studied the cuneiform-geometry implant considered with a 3-


dimensional model that had a mesh that was finer than in the models commonly found in the
literature. A mechanical model of an edentulous mandible was generated from computerized
tomography, with the implant placed in the left first premolar region. A 100-N axial load was
applied at the implant abutment, and the mandibular boundary conditions were modelled
considering the real geometry of its muscle supporting system. The cortical and trabecular bone
was assumed to be homogeneous, isotropic, and linearly elastic. It was concluded that geometry
showed a smooth stress pattern, with stress concentrated in the cervical region. The values,
however, were within the range of values found in the cortical layer far from the implant,
caused by the muscular action. No significant stress concentration was found in the apical area.

Geng J P et al (2001) : They reviewed the current status of Finite Element Analysis (FEA)
applications in implant dentistry and discusses findings from FEA studies in relation to the
bone–implant interface, the implant–prosthesis connection, and multiple-implant prostheses.

Kıvanç A et al (2001) : The study was conducted to evaluate the effect of staggered (offset,
tripodization) implant placement configuration and placement of wider-diameter implants in a
straight-line configuration in mandibular posterior edentulism. A mandibular Kennedy Class

REVIEW OF LITERATURE 25
II partially edentulous finite element model was constructed. Seven different partial fixed
prostheses supported by 3 implants were designed according to 2 main configurations: straight-
line or staggered implant placement. In 5 of the designs, implants with various diameters and
length were placed along a straight line. In the other 2 models, offset placement of the middle
implant buccally and lingually was simulated. A 400 N static load was applied perpendicular
to the buccal inclination of the buccal cusps on each unit. Tensile and compressive stress values
on cortical bone in the cervical region of the implants were evaluated. Lower stress values were
recorded for the configuration with wider implants placed in a straight line. Other
configurations, including staggered implant placement, produced similar stress values. Despite
the offset implant placement, the stresses were not decreased; however, straight placement of
wider implants may decrease bending moments.

Cehrili M et al (2002) : This study was conducted to evaluate the compatibility of three-
dimensional finite element stress analysis and in vitro strain gauge analysis in the measurement
of strains on a dental implant. It was done using two vertically placed implants embedded in a
poly (methyl methacrylate) model were used. Strain gauges were bonded to the cervical parts
of the implants, and seven cement-retained fixed partial dentures were fabricated. A three-
dimensional model of the strain gauge analysis model was constructed, and an additional model
in which human bone simulation was provided was also constructed. A static vertical load of
50 N was applied at certain locations to simulate centrally positioned axial and laterally
positioned axial loading for strain gauge analysis and three-dimensional finite element stress
analysis. It was seen that statistically significant increase in strain levels were recorded between
loading types in the strain gauge analysis. Strains obtained from strain gauge analysis were
higher than for three-dimensional finite element stress analysis. There was a remarkable
difference between the two finite element models under the conditions of laterally positioned
axial loading. Thus, concluding that there are differences regarding the quantification of strains
between strain gauge analysis and three-dimensional finite element stress analysis. However,
there is a mutual agreement and compatibility between three-dimensional finite element stress
analysis and in vitro strain gauge analysis on the determination of the quality of induced strains
under applied load.

REVIEW OF LITERATURE 26
Pierrisnard L et al (2002) : The study was to evaluate by finite element analysis the influence
of the design of 3 different dental implants on micromovements, cervical shearing stress
intensity, and stress distribution after occlusal loading. The first investigated implant was a
classical cylinder, the second was reinforced by two bicortical locking pins, and the third was
an expanding dental implant. The parameters analysed were the implant’s geometry, the quality
of the cancellous bone, and the orientation of occlusal loading. It was concluded that for the
cylindric implant, stresses were concentrated in the neck region; for the apical expansion
implant, stresses were evenly distributed from the neck to the apex of the implant. For the
locking pin implant, stresses around the neck were moderate and appeared concentrated around
the pins. Thus, initial stability of the pin implant was greater than that of the expanding implant,
but the expanding implant showed the most favourable stress distribution.

Ishigaki S et al (2003) : They conducted a study about to reveal the biomechanical stress
distribution in supporting bone around an implant and a natural tooth under chewing function.
Three-dimensional finite element models of the mandibular first molar and the titanium implant
both with the mandible in the molar region were constructed. The directions of displacement
constraints were determined according to the angles of the closing pathways of chopping type
and grinding type chewing patterns. The tooth model showed smooth stress distribution in the
supporting bone with low stress concentration around the neck of the tooth. The implant model
showed stress concentration in the supporting bone around the neck of the implant, especially
in the buccal area. The grinding type model of the implant showed higher tensile stress
concentration than the chopping type model at the lingual neck of the implant. The results of
this study suggested the importance of considering occlusion under chewing function for
understanding the biomechanics of oral implants.

Alkan I et al (2004) : They conducted a study to investigate the stress distribution of a


preloaded dental implant screws in three implant-to-abutment joint systems under simulated
occlusal forces using the 3-dimensional finite element analysis method:

(1) Branemark external hexagonal screw-retained abutment.


(2) ITI 8-degree Morse tapered cemented abutment.
(3) ITI 8-degree Morse tapered plus internal octagonal screw-retained abutment.

REVIEW OF LITERATURE 27
A thermal load and contact analysis method were used to simulate the preload resulting from
the manufacturers’ recommended torques in implant screw joint assemblies. The simulated
preloaded implants were then loaded with three simulated static occlusal loads (10 N
horizontal, 35 N vertical, 70 N oblique) on the crown position onto the implant complex.
Numeric and graphical results demonstrated that the stresses increased in both the abutment
and prosthetic screws in the finite element models after simulated horizontal loading. However,
when vertical and oblique static loads were applied, stresses decreased in the external
hexagonal and internal octagonal plus 8-degree Morse tapered abutment and prosthetic screws
with the exception of the prosthetic screw of ITI abutment after 70-N oblique loading. Stresses
increased in the ITI 8-degree Morse tapered cemented abutment after both vertical and oblique
loads. Thus, it was concluded although an increase or decrease was demonstrated for the
maximum calculated stress values in preloaded screws after occlusal loads, these maximum
stress values were well below the yield stress of both abutment and prosthetic screws of 2
implant systems tested. The results imply that the 3 implant-to-abutment joint systems tested
may not fail under the simulated occlusal forces.

Kitagawa T et al (2005) : They using finite element method (FEM), sought to investigate how
the thickness and Young's modulus of cortical bone influenced stress distribution in bone
surrounding a dental implant. The finite element implant-bone model consisted of a titanium
abutment, a titanium fixture, a gold alloy retaining screw, cancellous bone, and cortical bone.
The results showed that von Mises equivalent stress was at its maximum in the cortical bone
surrounding dental implant. Upon investigation, it was found that maximum von Mises
equivalent stress in bone decreased as cortical bone thickness increased. On the other hand,
maximum von Mises equivalent stress in bone increased as Young's modulus of cortical bone
increased. In conclusion, it was confirmed that von Mises equivalent stress was sensitive to the
thickness and Young's modulus of cortical bone.

Cruz M et al (2006) : They conducted a study about the biomechanical behaviour of an


osseointegrated dental implant plays an important role in its functional longevity inside the
bone. Studies of this aspect of dental implants by the finite element method are ongoing. In the
present study, a cuneiform-geometry implant was considered with a 3- dimensional model that
had a mesh that was finer than in the models commonly found in the literature. They made a

REVIEW OF LITERATURE 28
mechanical model of an edentulous mandible was generated from computerized tomography,
with the implant placed in the left first premolar region. A 100-N axial load was applied at the
implant abutment, and the mandibular boundary conditions were modelled considering the real
geometry of its muscle supporting system. The cortical and trabecular bone was assumed to be
homogeneous, isotropic, and linearly elastic. The results of the stress analysis were used to plot
global and detailed graphics of normal maximum (S1), minimum (S3), and von Mises stress
fields. The results obtained were analyzed and compared qualitatively with the literature. Thus
it was concluded the studied geometry showed a smooth stress pattern, with stress concentrated
in the cervical region. The values, however, were within the range of values found in the
cortical layer far from the implant, caused by the muscular action. No significant stress
concentration was found in the apical area.

Maeda Y et al (2007) : The purpose of this study was to examine the biomechanical advantages
of platform switching using three-dimensional finite element models. They used three
dimensional finite element models simulating an external hex implant (4 × 15 mm) and the
surrounding bone were constructed. One model was the simulation of a 4 mm diameter
abutment connection and the other was the simulation of a narrower 3.25 mm diameter
abutment connection, assuming a platform-switching configuration. The stress level in the
cervical bone area at the implant was greatly reduced when the narrow diameter abutment was
connected compared with the regular-sized one. It was concluded within the limitations of this
study, it was suggested that the platform switching configuration has the biomechanical
advantage of shifting the stress concentration area away from the cervical bone–implant
interface. It also has the disadvantage of increasing stress in the abutment or abutment screw.

Bergkvist G et al (2008) : Their study was conducted using the finite element method (FEM)
to simulate stresses induced in bone tissue surrounding uncoupled and splinted implants in the
maxilla because of bite force loading, and to determine whether the differences in these stress
levels are related to differences in observed bone losses associated with the two healing
methods. It was seen that the stress levels in bone tissue surrounding splinted implants were
markedly lower than stress levels surrounding uncoupled implants by a factor of nearly 9.

REVIEW OF LITERATURE 29
Bellini C M et al (2009) : They studied , the stress patterns induced in cortical bone by three
distinct implant-supported prosthetic designs , using finite element analysis. They constructed
two models consisted of a prosthesis supported by four implants, the distal two of which were
tilted, with different cantilever lengths (5 mm and 15 mm). The third design consisted of a
prosthesis supported by five conventionally placed implants and a 15-mm cantilever. They
results showed in the tilted model with 5-mm cantilever and in the non-tilted model, the
maximum value of compressive stress (–18 MPa) was found near the cervical area of the distal
implant. Higher values for compressive stress were predicted near the cervical area of the distal
implant in the tilted model with a 15-mm cantilever, as compared to the tilted model with the
5-mm cantilever. For the tilted model with the 5-mm cantilever, peak values of tensile stress
were predicted near the cervical area of both the distal (1.25 MPa) and the mesial implants (2.5
MPa). For the non-tilted model, the peak value was found near the cervical area of the in-
between implant (5 MPa). For the tilted model with 15-mm cantilever, tensile stress values
were higher than in the tilted model with 5-mm cantilever. They concluded that no significant
difference in stress patterns between the tilted 5-mm and the non-tilted 15-mm configuration
was predicted. The tilted configuration with a 15-mm cantilever was found to induce higher
stress values than the tilted configuration with a 5-mm cantilever

Djebbar N et al (2010) : In this study the finite element method is used to compute the
distribution of stresses in dental prosthesis. The stress analysis is particularly focused at the
interface bone–implant in different positions: distal zone, medial zone of these components.
The effects of the intensity and the direction of loading on the stress were concluded to be very
minimal.

Jose H R et al (2010) : They conducted a study in which many clinical variations present in
implant‐supported prosthesis were analysed by 3‐D finite element method. A geometrical
model representing the anterior segment of a human mandible treated with 5 implants
supporting a framework was created to perform the tests. The variables introduced in the
computer model were cantilever length, elastic modulus of cancellous bone, abutment length,
implant length, and framework alloy (AgPd or CoCr). The computer was programmed with
physical properties of the materials as derived from the literature, and a 100N vertical load was
used to simulate the occlusal force. Images with the fringes of stress were obtained and the

REVIEW OF LITERATURE 30
maximum stress at each site was plotted in graphs for comparison. It was seen that stresses
clustered at the elements closest to the loading point. Stress increase was found to be
proportional to the increase in cantilever length and inversely proportional to the increase in
the elastic modulus of cancellous bone. Increasing the abutment length resulted in a decrease
of stress on implants and framework. Stress decrease could not be demonstrated with implants
longer than 13mm. A stiffer framework may allow better stress distribution. Thus, it was
concluded that the relative physical properties of the many materials involved in an implant‐
supported prosthesis system affect the way stresses are distributed.

Kim K S et al (2011) : The purpose of this study was to examine photoelasticity effect on the
inclination of the two distal implants according to the All-on-Four concept on the stress
distribution within the supporting structure. They constructed two photoelastic models of a
human edentulous mandible were fabricated. Each model had four screw-type implants
embedded in the interforaminal area. The two distal implants were placed axially in one model
and tilted 30 degrees distally in the other model. Two cantilevered acrylic resin prostheses,
which used angulated abutments for the distal tilted implants and straight abutments for the
axial implants, were fabricated and delivered. Vertical loads of 13 kg were applied at three
loading points on the prosthesis: the central fossa of the first molar, the distal fossa of the first
premolar, and the distal fossa of the second premolar. Stresses that developed in the supporting
structure were monitored photoelastically and recorded photographically. It was seen all
cantilever loadings concentrated the stresses at the distal crest of the distal implant sites in both
models, the posterior tilting of distal implants splinted in a full-arch fixed prosthesis did not
increase the stresses in bone around the distal implants versus the axial-implant model. Thus,
it was concluded within the limitations of this photoelastic stress analysis, the use of tilted
implants reduced the maximum stress in the distal crestal bone of the distal implant by
approximately 17% relative to the axial implants

Ozkir S E, Terzioglu H E (2012) : They using the Photo-elastic Stress Analysis Method (PSA),
concluded that screw cylinder implants with micro threads on the implant neck are useful at
stress distribution. Stepped cylinder implants have better stress distribution properties when
placed inclined. Screw cylinder implants have acceptable stress distribution properties. Root
form implants have greater stress concentrations than the other types when they transfer stress.

REVIEW OF LITERATURE 31
Woo Taek Lee et al (2012) : They conducted a study to evaluate the fatigue limits of PEEK
and the effects of the low elastic modulus PEEK in relation to existing dental implants.
Compressive loading tests were performed with glass fibre‐reinforced PEEK (GFR‐PEEK),
carbon fibre‐reinforced PEEK (CFR‐PEEK), and titanium rods. Among these tests, GFR‐
PEEK fatigue tests were performed according to ISO 14801. For the finite element analysis,
three‐dimensional models of dental implants and bone were constructed. The implants in the
test groups were coated with a 0.5‐mm thick and 5‐mm long PEEK layer on the upper intrabony
area. The strain energy densities (SED) were calculated, and the bone resorption was predicted.
The fatigue limits of GFR‐PEEK were 310 N and were higher than the static compressive
strength of GFR‐PEEK. The bone around PEEK‐coated implants showed higher levels of SED
than the bone in direct contact with the implants, and the wider diameter and stiffer implants
showed lower levels of SED. The compressive strength of the GFR‐PEEK and CFR‐PEEK
implants ranged within the bite force of the anterior and posterior dentitions, respectively, and
the PEEK implants showed adequate fatigue limits for replacing the anterior teeth. Dental
implants with PEEK coatings and PEEK implants may reduce stress shielding effects.

Cardoso M V , Chaudhari A et al (2013) : They compared the clinical performance of two


dental implant types possessing a different macro-design in the in vivo pig model. It was found
that a decreased thread pitch can provide a positive contribution to BIC and to the implant
stability, via an increased surface area and a better force distribution. Consideration of specific
implant macro-design features should be made relative to other factors such as the biological
and mechanical microenvironment.

Hansson S et al (2014) : A study combining the three dimensional and asymmetric finite
element analysis was done to measure the effect of the peak interfacial stress on providing the
axial loaded mandibular dental implant with retention elements all the way upto the crestal
bone. The interpretation of this was all the subjected stresses increased the capacity of the
implant to carry off axial loads under measured control.

Gehrke S A, (2015) : The purpose of their study was to assess implant stability in relation to
implant design (conical vs. semi conical and wide-pitch vs narrow pitch) using resonance

REVIEW OF LITERATURE 32
frequency analysis. It was concluded that Macro design includes thread pitch, body shape, and
thread design, while micro design essentially regards the surface morphology. The primary
stability of dental implants is highly dependent on surgical technique and bone features at the
implant site. Primary implant stability has long been considered a fundamental predictor for
successful osseointegration. Biomechanical and finite element analysis studies showed that
maximum effective stress decreased as screw pitch decreased and implant length increased.
Wider thread pitch is related to better implant primary stability, providing a higher mechanical
interlocking with bone tissue.

Yadav P, Shetty P et al (2016) : They conducted a study to investigate about the stress and
strain fields around osseointegrated dental implants that maybe affected by a number of
biomechanical factors, including the type of loading, material properties of the implant and the
prosthesis, implant geometry, surface structure, quality and quantity of the surrounding bone,
and the nature of the bone–implant interface.

Andrade C L et al (2017) : The aim of their study was to evaluate the influence of implant
macro-design when using different types of collar an and thread design on stress distribution
in a maxillary bone site. The von Misses stresses and the tensile stress were measured using
three dimensional finite element analysis and ANOVA was done. It was concluded that the
stress/strain pattern were influenced by collar design in the implant and cortical bone and by
thread design in trabecular bone. Micro-threads and triangular geometry thread shapes were
presented with improved biomechanical behaviour in posterior maxilla bone when compared
to smooth collar design and trapezoidal square shaped threads

Jaros O A et al (2018) : This study assessed by using finite element analysis, the
biomechanical behaviour of an implant system using the All-on-Four technique with nickel–
chromium (M1) and polyether ether ketone (PEEK) bars (M2). Data were analysed according
to system's areas of action: peri-implant bone, implant, intermediates, intermediates’ screws,
prostheses’ screws, and bars. Largest peak stress was shown in M2.

REVIEW OF LITERATURE 33
TOOTH VERSUS IMPLANT SUPPORT SYSTEMS
TOOTH IMPLANT
1. PRESENCE OF DIRECT BONE IMPLANT
PERIODONTIUM

a) Shock absorber. Higher impact force.


b) Distribution of force around tooth. Force primarily to crest.
c) Tooth mobility can be related to Implant is always rigid.
force.
d) Mobility dissipates lateral force. Lateral force increases strain to
bone.
e) Fremitus related to force. No Fremitus.

f) Radiographic changes related to Radiographic changes at crest


force (reversible). (not reversible).

2. BIOMECHANICAL DESIGN IMPLANT DESIGN

a) Longer force duration Short force duration (increased


(decreases impulse of force). force impulse).
b) Cross section related to Round cross-section and designed
direction and amount of stress. for surgery.
c) Elastic modulus similar to that Elastic modulus 5 to 10 times that
of cortical bone. of cortical bone.
d) Diameter related to force Diameter related to existing bone.
magnitude.

3. SENSORY NERVE COMPLEX NO SENSORY NERVES THUS NO


IN AND AROUND TOOTH PRECURSOR SIGN OF SLIGHT
OCCLUSAL TRAUMA

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 34


a) Occlusal trauma induces a) Occlusal awareness of two to five
hyperemia and leads to cold times less (higher maximum bite
sensitivity. force functional).
b) Proprioception (reduced maximum Functional bite force four times
bite force). higher.

4. OCCLUSAL MATERIAL OCCLUSAL MATERIAL


ENAMEL VARIABLE

a) Enamel wear, stress lines, a) Not much effect of force.


abfraction and pits

5. SURROUNDING BONE IS SURROUNDING BONE IS


CORTICAL TRABECULAR
a) Resistant to change. a) Conducive to change
b) High strength. b) Reduced strength

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 35


Fig : 4.1

DIFFERENCES BETWEEN NATURAL TEETH AND IMPLANTS :

Differences between natural tooth and endo-osseous dental implants under occlusal
loading are summarized in the above table. The basic difference between natural teeth and
endo-osseous dental implants is that a natural tooth has a support design that reduces the forces
to the surrounding crest of bone compared to the same region around an implant. A natural
tooth is suspended by the periodontal ligament while an endo-osseous dental implant is in direct
contact with the bone through osseointegration. The periodontal ligament absorbs shocks and
distributes occlusal stresses away along the axis of natural teeth. However, an endo-osseous
dental implant connected to the bone by osseointegration lacks those advantages of the
periodontal ligament. Teeth in natural dentition are retained by periodontal tissues that are
uniquely innervated and structured. When natural teeth are lost, both occlusion and attachment
with its proprioceptive feedback mechanism are lost. When loaded, the movement patterns of
natural teeth begin with the primary phase of periodontal compliance that is primarily non-
linear and complex, followed by the secondary movement phase which occurs with
engagement of the alveolar bone. In contrast, the movement of an implant under loading is

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 36


dependent on linear and elastic deformation of the bone. The periodontal ligament in a natural
tooth can produce differences in force adaptation compared with osseointegrated implants due
to its shock-absorbing and stress-distributing functions.

Non-vertical forces on natural teeth during function affect only the teeth involved and
are usually tolerated, whereas in implants, the effect involves the crest of the bone, which is
usually traumatic to the supporting structures. A lateral force on a healthy natural tooth is
rapidly dissipated away from the crest of bone toward the apex of the tooth due to the natural
tooth rapidly moving 56-108 μm and rotating around the apical 1/3rd of the root. On the other
hand, movement of an implant occurs gradually, reaching up to about 10-50 μm under a similar
lateral force. Thus greater forces are concentrated on the crest of the surrounding bone of dental
implant in the absence of rotation. Under similar lateral loads, an implant does not pivot as
much as a tooth toward the apex, but instead concentrates greater forces at the crest of the
surrounding bone. Therefore, if an initial load of equal magnitude and direction is placed on
both an implant and natural tooth, the implant must be protected.

Malocclusion of natural teeth may be uneventful for years. However it may evokes a
traumatic response and involves the crest of the surrounding bone. Richter’s et al (1998)
studied and reported that a transverse load and clenching at centric contacts resulted in the
highest stresses in the crestal bone of dental implants. Misch (1989) suggested that gradient
loading to accommodate the disadvantageous kinetics associated with dental implants in
patients with a poor bone quality condition. In natural teeth, proprioception gives the
neuromuscular system control during function. This makes it possible for a person to avoid
prematurity and interferences, and to establish a stable habitual occlusion away from a centric
relation. With dental implants, no such feedback signal system is present, and the mandible's
function will end its chewing stroke in the most favourable kinesiologic position, which is very
close to a centric relation. If cusps interfere or prematurity exist as the mandible returns to this
position, crest bone loss will occur. The presence or absence of the Periodontal Ligament’s
function makes a remarkable difference in detecting the early phase of occlusal forces between
teeth and implants. Because periodontal mechanoreceptors in natural teeth provide
proprioception and early detection of occlusal forces and interferences, the bite forces used in
mastication and parafunction are not as strong due to fine motor control of the mandible.

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 37


Trulsson et al (2005) also reported that implant denture patients split a peanut with a
force 4-fold greater than that of the natural dentition group and pointed out that a lack of
proprioception can lead to a heavier bite force in implant patients. Mericske-Stern et al (1994)
measured the oral tactile sensibility with test steel foil. The detection threshold of minimal
pressure was significantly higher with implants than with natural teeth (3.2 vs. 2.6 steel foil
sheets). Jacobs and Van Steinberg et al (2001) evaluated the occlusal awareness and found
that interference perceptions of natural teeth, implants with opposing teeth, and implants
opposing implants were approximately 20, 48, and 64 μm, respectively. Hammerle et al (1996)
also concluded that the mean threshold value of tactile perception for implants (100.6 g)) was
9-fold higher than that of natural teeth (11.5 g).

Clinical evidence of occlusal trauma on teeth includes an overall thickening of the


periodontal membrane, tooth mobility, and increased radiopacity and thickness of the
cribriform plate around a tooth, as observed on radiographs and not just localized at the crest.
A tooth can show clinical signs of increased stresses such as enamel wear facets, cervical
abfraction, fremitus, and pain. When implants are subjected to repeated excessive occlusal
loads, no generalized radiographic signs are apparent around an implant, except at the crestal
region, which demonstrates bone loss but may be misdiagnosed as peri-implant disease due to
bacteria and pathogens. Implant components rarely show clinical signs other than fatigue
fractures (screw loosening or fracture, abutment or prostheses fracture, and implant fracture).
Dentists can replace a natural tooth with an artificial one but not its attachment, which presents
a new problem, and it seems logical that some changes must be made. The above differences
necessitate consideration of occlusion for dental implants as a special problem with different
requirements if they are to function efficiently with the least amount of trauma to supporting
tissues. The physiological differences between a natural tooth and endosseous dental implant,
it can be concluded that osseointegrated implants without periodontal receptors would be more
susceptible to occlusal overloading. Because the load-sharing ability, adaptation to occlusal
forces, and mechanoperception are significantly reduced in dental implants, there are
differences in occlusal considerations between natural teeth and implants.

BITE FORCES AND IN VIVO LOADING OF IMPLANTS BITE FORCES IN THE


NORMAL DENTITION:

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 38


In the normal dentition without implants, mean maximal vertical (axial) bite force
magnitudes in humans can be 469 ± 85 N at the region of the canines, 583 ± 99 N at the second
premolar region, and 723 ± 138 N at the second molar. A vertical force is defined as the force
component acting perpendicular to the occlusal plane. In the incisal region, the direction of
maximum incisal bite force is about 12 degrees to the frontal plane, which suggests that the
lateral components of force on an anterior implant could be appreciable.

BITE FORCES AFTER IMPLANT TREATMENT:

It has been suggested that the general features of mastication in patients with normal
and implant-restored dentitions are approximately the same. However, Carr and Laney (1987)
reported a significant improvement in both maximum and mean biting forces in a group of
fourteen patients who started with conventional complete dentures but then received
mandibular tissue-integrated prostheses opposing a complete maxillary denture.

Mericske-Stern et al (1994) explored maximum occlusal force and oral tactile


sensibility in a group of partially edentulous patients restored with ITI implants supporting
group of fully dentate subjects with healthy natural teeth. They found the highest maximal
occlusal force in fully dentate subjects on the second premolars (mean 450 N). For fixed
prostheses supported by implants, the average value of maximal occlusal force was lower—
about 200 N for first premolars and molars and 300 N for second molars. In their data, the
range of forces measured in the various subjects included values as large as 1,100 N in the
molar region.

IN VIVO MEASUREMENTS OF BITE FORCES AND MOMENTS ON IMPLANTS:

Implants used for single tooth replacements, the in vivo forces acting are sought to
replicate the forces exerted on natural teeth. This is expected because in both cases, the biting
would be delivered to single, stand-alone crowns. However, factors such as the width of the

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 39


crown occlusal table, the height of the abutment above the bone level, and the angulations of
the implant with respect to the occlusal plane will affect the value of the moment on the implant.

Under some conditions, leverage effects exist because of geometric factors relating to
restorations linking the implants, such as the existence of distal cantilevers in a full-arch
restoration. Such factors cause the implants to be subjected to increased bending moments as
well as axial forces that can be tensile and compressive.

Moreover, when biting occurred at the cantilever location of a prosthesis supported by


five implants, axial forces on some of the implants were as much as twice the biting force on
the prosthesis, a result that can be predicted from theoretical models.

Rangert et al (1997) used the same methods for in vivo measurements of the vertical
load distribution and bending moments on a 3-unit prosthesis supported by a natural tooth and
a single Brånemark implant. In those five patients, they demonstrated that the vertical loads
were distributed between the natural tooth and the implant.

Also, axial forces on an implant could be tensile or compressive, depending on their


location when supporting 3-unit prosthesis, and that the axial loading on an implant could
sometimes be twice as large as the bite force on the prosthesis because of cantilever extensions.

Richter et al (1998) used a strain-gauged abutment to measure in vivo horizontal


bending moments on IMZ implants in the molar area in humans. They reported that maximum
bending moments and the corresponding transverse force when the patients chewed various
types of food, including sticky confections (jelly beans), sausage, carrots, and crackers. The
magnitudes of the bending moments ranged from less than 10 N-cm to slightly more than 20
N-cm. The corresponding transverse forces in buccal and oral directions were always less than
30 N.

TOOTH VERSUS IMPLANT SUPPORT SYSTEMS 40


COMPONENTS OF FORCE

CHARACTER OF FORCES APPLIED TO DENTAL IMPLANTS

Fig 5.1

• Forces may be described by magnitude, duration, type and magnification factors.


• Forces acting on dental implants are ‘Vector Quantities’ as they possess magnitude and
direction.
• There is a significant influence of load direction on implant longevity.
• A force applied to a dental implant rarely is directed absolutely longitudinally along a
single axis. In fact three dominant clinical axes exist in implant dentistry. Mesiodistal,
Faciolingual and Occlusoapical and they commonly result from single occlusal contact.
• The process by which three-dimensional forces are broken down into their component
parts is referred to as Vector Resolution and may be used routinely in clinical practice for
better understanding and incorporating features which will produce enhanced implant
longevity.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 41


• Along with the direction of force it Is also important to specify the point of action of a
vector.

Fig 5.2

Diagram of a tooth loaded by 44.5 N force acting along the line of action which is
perpendicular to the surface of the tooth at A and not parallel to the long axis of a tooth

• Vectors are usually written in bold faces (F) or with an arrow above F and magnitude is
written as simply F.
• Suppose a 44.5 N force arises due to point contact or chewing at Point A on the crown
supported by single implant and this force is not directed parallel to the direction of long
axis of implant.
• So, the force will have two parts, one which will act parallel and another which will be
perpendicular to the axis of implant.
• It must be noted that an implant or some part of it can fracture if the perpendicular
component becomes too large.
• Thus to analyze this problems it is easier to resolve the force vector into components
along the directions of interest.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 42


• Thus a co-ordinate system with x, y and z axes at right angles to one another and with 2-
axis parallel to long axis of implant is selected.
• Then considering the angles that the force vector makes with three co-ordinate axes, it is
possible to resolve the force into its three components. i.e., Fx, Fy and Fz

Fig 5.3

So, by using mathematical formulae

Fx = F cos x Fy = cos xy Fz = cos z

When F = scalar magnitude of force = 44.5 N

x, y and z are angles between force vectors x-y and z axes respectively.

F can be calculated from (magnitude of force vector) Fx , Fy , Fz

As F = Fx , Fy , Fz

• Another relation between the angles x, y and z is


cos2 x2 + cos2 y + cos2 z = 1

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 43


• So if we know x and y its possible to calculate z and then calculating values of Fx, Fy,
Fz similarly.
• So with the above analysis it is clear that there are lateral as well as vertical force
components acting at the same time on a tooth or implant but it is true that the largest
component is vertical.
• During grinding of the teeth, the lateral component may be largest.

• Vector Addition :

If more than one force is acting on some object then the resulting force is the vectorial
sum of all the forces acting on the body and the Force Resultant (FR) is formed from a
vector sum of F1 + F2 + F3.

Fig 5.4

• Moment / Torque :
- Basically its an action which tend to rotate a body.

S.I. unit – N.m.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 44


Fig : 5.5
• Deformation and strain :

Fig 5.6

o A load applied to implant can cause deformation of implant and surrounding tissue.

o Surrounding tissue can react to it by remodelling.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 45


o Deformation and stiffness characteristics of implant material may influence interfacial
tissue, and interfere with longevity.

o Concept of strain is believed to be a key mediator of bone activity.

STRESS - STRAIN RELATIONSHIP

Fig 5.7

o If any elastic body is subjected experimentally to an applied load, a load-versus-

deformation curve can be generated.

o So if we divide the load (Force) by the surface area over which they act and the charge

in length by the original length produces a classic stress-strain curve.

o Such curve provides for the prediction of how much strain will be experienced in a

given material under the action of an applied load.

o Linear portion of this curve is referred to as modulus of elasticity.

o Closer the modulus of elasticity of the implant to surrounding biologic tissues, the lesser

the chances of relative motion at bone-implant interface.

o Cortical bone is 5 times more flexible than titanium.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 46


o So as stress increases relative stiffness difference increases and thus relative motion

increases and interface is more affected.

o Other way round, viscoelastic bone can stay in contact with more rigid titanium more

predictably when the stress is low.

o Once an implant system is selected, the only way for a clinician to control the strain is

by controlling applied stress or change the density of bone around the implant.

o The density of bone is not only related to strength but also to the stiffness (Modulus Of

Elasticity).

o So, the relative difference in stiffness is less for Titanium or its alloy and D1 bone than

compared with Dfour bone with the same.

o Thus reducing the stress in such softer bone is to reduce the resultant tissue strains

resulting from the elastic differences because softer bone exhibits a lower ultimate

strength.

o Stress and strain can be related by a mathematical equation according to Hook’s law as,
€ = E

€ = Stress (Pascal / Pounds per sq. inch)

E = Modulus of elasticity.

 = Strain (unitless)

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 47


TYPICAL AND MAXIMUM BITING FORCES

Fig 5.8

o Normal human (No prosthesis) can typically exert axial components of biting forces in

the range of 100-2400 N (27-550 lbs).

o Axial components on natural teeth tend to be larger as me moves distally in the mouth.

It can be explained with Example of mandible as a class 2 lever.

o Fulcrum is at the condyle (C); while the two major muscle forces M1 and M2 act nearer

to the fulcrum than biting force (F). This class 3 lever has mechanical advantage of less

than 1 and bite force will be larger if it acts nearer to the fulcrum i.e., molars TMJ.

o Typically lateral components of force were about 20 N in patients with prosthesis in the

first mandibular molar region.

o The net chewing time per meal – 450 Sec. If chewing frequency is 1.5-2 per second

with a 0.3 sec duration of tooth contact per chewing stroke. There will be about 9 min

/ day chewing forces will act on teeth. If other activities such as swallowing are

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 48


considered the time might increase to 17.5 min/day. Still which might further be

increased if there is any parafunction associated.

o These estimates provide a useful indication of minimum time/day that teeth (implants)

are load bearing/ subjected to load due to mastication and other related events.

PREDICTING FORCES ON ORAL IMPLANTS :

o Assuming that the biting force on a prosthesis are known, it is not always a simple

individual supporting abutments. As thee forces will not be the same as exerted on the

prosthesis.

o The problem of calculating the forces on individual abutment especially in cases with

more than two implants supporting a prosthesis.

o Many other complicating factors can be involved as :

1) The nature of mastication :

- Frequency of biting,

- Strength of biting,

- Sequence of chewing cycle,

- Mandibular movements.

2) The nature of prosthesis :

- Full or partial dentures,

- Tissue-supported versus implant –supported prosthesis,

- Number and location of implants and teeth,

- Angulation of implants.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 49


3) The biomechanical properties of the structures and materials comprising the

bridge or prosthesis, implants and bone. eg : elastic moduli, structural stiffness,

nature of the connection between implant and bridge and deformability of

mandible or maxilla.

We will see the methods to account some of these factors with the help of models but one has
to consider its limitations with respect to a given clinical situation.

• The diagram shows a method for predicting the forces on two implants supporting a
cantilever portion of a prosthesis. (Rangert Model 1989).
• A downwards force P acts at the end of a bridge with a cantilever section of length a (i.e.
distance between line of action and the nearest implant).

Fig: 5.9

• Bridge is assumed to be a rigid beam supported by two implants separated by distance b.

• F1 and F2 are forces that implants exert on the beam. As beam is in static equilibrium;

according to Newton’s laws this means that the sum of the forces and the sum of the

moments on the beam are both zero.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 50


Fy = 0; - F1 + F2 – P = 0  MQ = 0 ; - F1b + aP =0

Fy = summation of forces in y axis MQ = summation of moments about Q

• Solutions for F1 and F2 from the previous equations can be derived are :

F2 = (1+a/b) P F1 = (a/b) P

• Although the bridge is loaded by biting force P the implants are loaded by forces whose

magnitude is greater than P depending on ratio a/b.

• In most clinical practice it can be considered a/b = 2 so, forces on the implants are :- 3P

and 2 P.

According to Newton’s 3rd law of motion – For every action there is equal and opposite

reaction. In any ease the forces F1 and F2 do not act in the same direction.

• Implant No.2 nearest to the point of force, experiences a compressive load, tending to push

it into the bone.

• Implant No. 1 experience a tensile load, tending to pull it out of the bone.

Fig 5.10

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 51


• Four maxillary implants supporting a maxillary framework in the Branemark system.

• The dist. a and b can be measured chairside. Forces on the two implants nearest to the

applied force P can be obtained as like in the previous scenario.

• Limitation –

1) Does not predict forces on all four implants.

2) Thus Rangert model will overestimate the loads on the two implants nearest to the

applied load P.

• The abutment loadings cannot be obtained using only the theory of rigid body static but it

is possible to solve the problem if information about the mechanical properties of the

bridge, abutments and interfacial tissues is available.

SKALAK MODEL (1985) was the first solution of this sort of problem.

• It can predict the vertical and horizontal force components on implants supporting a bridge

subjected to vertical and horizontal loadings.

• It is assumed that bridge and bone are rigid but the implants and / or their connections to

bridge and / or bone are elastic.

• This model predicts that, a purely vertical force on the prosthesis is counter balanced by a

distribution of purely vertical forces among the N number of supporting abutments.

Similarly, for a horizontal load on prosthesis, the model predicts that there will be a counter

balancing distribution of horizontal forces among the N number of abutments.

• Thus there will be both vertical and horizontal force components on each of the implants.

• Four or six implants symmetrically distributed about the midline of a mandible.

• Arc of distribution is 112.5 (distance between two mental foramina) (Approx.) and radius

of mandible equal to 22.5 mm.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 52


• For six implant case, the most distal implants nearest the load (i.e., No. 1 and 2) experience

compressive forces (negative values) similarly implant no. 6.

• On the other side will experience meanwhile, the three anterior implants.

No. 3, 4, 5 experiences tensile forces (positive values) thus preventing bridge tipping distally

and to the side.

- Applied vertical load magnitude is of 30 N but the loads on the implants are less than

30N except for that on implant No. 1 (nearest to loading patient).

Consider the four implants distributed over the same are as six implants (112.50). The results

show that the magnitudes of forces on the most distal implants are similar in both cases. This

means that there is only a slight different between using four implants and six implants to

support a prosthesis, when four implants are spaced out over a the same area as that of six

implants.

F< 30 N Magnitde of force is |||

Fig : 5.11 and 5.12

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 53


• Note that the inter-implant spacing in the four implant case is larger than in six implant

scenario this is to compensates for fewer implants.

• Now, consider a new arrangement of four implants created by removing two most distal

implants from six implant case, keeping inter implant spacing the same. In this case forces

on the four remaining implants become much larger than in the original six implant case.

Condition will be further worsened if four implants placed in a straight line across the

anterior of mandibular.

• In terms of Rangert model, the ratio a/b is very large as b is very small.

• All above models do not consider the effect of implant angulations.

Same diagram of cantilever prosthesis supported by two implants with are of the implants

at 300 inclination to the vertical. The only different that the force on implant no. 1 causes an

off-axis loading of the implant. This situation can lead to problems with implant or the bone

perhaps both and cannot be solved by Skalak or Rangert models.

Fig 5.13

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 54


• More recent analysis of forces on implants supporting bridgework have been accomplished

by a powerful computer method called Finite Element Modelling (FEM) or Analysis (FEA)

thus properties of the prosthesis, position and angulations of implants, properties of

interfacial bone can be accounted to FE models.

• The SKALAK MODEL assumes that prosthesis is infinitely rigid, which is obviously not

quite accurate.

• The acrylic and metal alloy bridge show a degree of flexibility, which has effect of

concentrating forces on those implants nearest to the loading point.

• Things become complicated when implants in system do not have equal stiffness. In such

situation results cannot be generalized as stiffest implant will generally take most of the

load.

• So, one should have in depth knowledge of tooth and implant stiffness.

Fig : 5.14 (SKALAK MODEL SHOWING RIGID PARAMETERS)

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 55


STIFFNESS OF TEETH AND IMPLANT :

If one considers a prosthesis supported by both teeth and implants, the difficulty is that

the teeth and implants do not have same characteristics for example mobility and in such

case neither Rangert nor Skalak model can specifically deal with differing mobility

among abutments.

SPECIFIC PROBLEMS ASSOCIATED WILL BE :

1) Teeth and implant can be displaced in any direction.

2) When a constant force is applied to a tooth or implant, the displacement of the tooth or

implant may increase slowly with time, this phenomenon is called creep. Creep in

relation with implants is not significant till they have fibrous tissue around.

3) Intrusive tooth displacement is not always linear with intrusive force. Usually it’s a

bilinear relationship.

Most implants in bone produce a net stiffness greater than for natural teeth.

Angled abutments also result in development of dangerous transverse force components

under occlusal loads in the direction of the angled abutment. Implants should be placed surgi-

cally to provide for mechanical loading down the long axis of the implant body to the maximum

extent possible. Angled abutments are used to improve aesthetics or the path of insertion of a

restoration, not to determine the direction of load.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 56


Fig : 5.15

Occlusion serves as the primary determinant in establishing load direction. The position

of occlusal contacts on the prosthesis directly influences the type of force components distrib-

uted throughout the implant system. The dentist should visualize each occlusal contact on an

implant restoration in its component parts. Consider the example of a restored dental implant

subjected to a premature contact during occlusion. When the contact is broken down into its

component parts directed along the three clinical loading axes, a large, potentially dangerous

lateral component is observed. Occlusal adjustments consistent with implant-protective

occlusion to eliminate the premature contact minimize the development of such dangerous load

components.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 57


FORCE MAGNITUDE:

A surgical placement resulting in extreme angulation of the implant, and/or a patient

exhibiting parafunctional habits will likely exceed the capability of any dental implant design

to withstand physiologic loads. Cantilevers and crown heights are levers and therefore force

magnifiers.

Careful treatment planning with special attention to the use of multiple implants to

increase functional surface area is indicated when a clinical case presents the challenge of force

magnifiers. A magnifier of force around an individual implant is also related to the density of

bone. Since density is directly related to bone strength and D4 bone is 10 times weaker than D1

bone, the effect of this resultant force is magnified as to its clinical result when placed upon

softer bone types.

As a result, considerable effort is made to decrease the effects of compromised bone

density, including implant design, size, coatings, and progressive loading to decrease stress

and/or increase bone strength.

COMPONENTS OF FORCE APPLIED TO DENTAL IMPLANTS 58


BIOMECHANICAL RESPONSE AT THE DENTAL
IMPLANT-TISSURE INTERFACE

BONE RESPONSE TO MECHANICAL LOADS :

The implant-to-tissue interface is an extremely dynamic region of interaction.


This interface completely changes character as it goes from its genesis ( from the
placement of the implant into the prepared bony site) to its maturity (healed condition).
The biomechanical environment plays an immediate role in the quality and
compositional outcome of the new interface. For example, extensive research shows that
if the implant is stable in the bone at the time of placement, the interface is more likely
to result in osteointegration. Relative movement (or micromotion) between the implant
and the bone at the time of placement is more likely to favour the development of a fibro
osseous interface. The healing stage of the interface, however, is only the beginning of
its dynamic nature. Functional loading of the implant brings additional biomechanical
influences that greatly affect the composition of this junction.

It has been proven that bone responds to both hormonal and biomechanical
(functional loading) regulation. These two regulating mechanisms are often in opposition
to each other. It has been theorized that the actual strain that is perceived by the bone
tissue initiates a chain of events that result in a biologic response. For tissues the strain
to influence bone adaptation at the bone-to-implant interface, it must elicit some sort of
a chemical or biologic response. The current hypothesis is that bone cells in conjunction
with the extracellular matrix comprise the strain-sensitive environment and that each
plays a vital role in the mediation of the interface. Based on this rationale, the objective
of a good implant design would be to establish and maintain a strain environment within
the host bone tissue and at the interface that favours osteointegration of the implant.

BIOMECHANICALLY-BASED BONE REMODELING THEORIES :

The desire to optimize the effects of strain at the bone-to- implant interface to
encourage osteointegration was largely fostered by biomechanical-based bone remodeling
theories.

BIOMECHANICAL RESPONSE 59
In Meier (1887) described the systematic structure of trabecular bone in the femoral
head in his book ‘Die Architektur Der Spongiosa ‘ (Cancellous Bone Architecture).

In 1888, the swiss physicist Kulmann found structural similarities between the
sketches of femur heads published by Meier and the course of tension trajectories in bent
girder beans Kulmann had calculated.

In 1892, Wolff described these events as a law of nature and stated that the trabecular
bone will place or displace itself in relationship to the functional pressures.

In 1895, Roux suggested that the tissue changes to loading were a result of a cellular
regulation process.

H.M. Frost (1996) proposed the theory of the Mechanostat. He postulated that bone
mass is a direct result of the mechanical usage of the skeleton. This agrees with Wolff’s
law, which in summary states that ‘form follows function’. Frost established a mechanical
adaptation chart relating trivial loading, physiologic loading, overloading, and pathologic
loading zones to ranges of micro strain. His studies showed that strains in the range of 50
to 1500 micro-strain stimulated increases in cortical bone mass until the strains were
reduced to the threshold range or minimum effective strain. This process of the mechanostat
would effectively switch the bone modelling on and off.

Bone may reduce strains by bone formation or resorption, and by changing modulus
of elasticity or stiffness by changing mineral content. Necrosis of bone cells appears to
determine the mineral content upper equilibrium level. Cell destruction can be observed
when stresses exceed 6.9 X 10 N/mm2, whereas a stress of 2.48 x 10 N/mm2 will cause an
increase in bone growth.

Cowin (2004) proposed potential mechanisms by which bone cells sense mechanical
load. Tissue-level strains were regarded as macroscopic strains averaged over a significant
volume of bone tissue, whereas cell-level strains were defined as highly localized strains
at the cell lacunae level. He also suggested that cell-level strains were almost tenfold greater
than tissue-level strains based on the confines and geometry of the lacunar shape. The
proposed mechanisms included membrane deformation (action potential), intracellular
action (passage of secondary messengers), and extracellular action (streaming potentials).

BIOMECHANICAL RESPONSE 60
Hasegawa and Binderman (1985) found that when bone cells were mechanically
distorted in cell culture, DNA synthesis was increased by 64% within 2 hours, or
phospholipase A2; was activated.

Numerous in vitro techniques have been used to stimulate mechanical loading at the
cellular level. Duncan and Turner (1995) developed schematic drawings depicting
variations in cellular deformation resulting from different in vitro loading schemes. The
most commonly reported loading methods include hypotonic swelling, hydrostatic
pressure, uniaxial stretch, biaxial stretch, and fluid shear stress.

It has also been shown that dynamic or cyclic loading is necessary to cause a
significant metabolic change to occur in the bone cell population both in vivo and in vitro.
The greater the rate of change of applied strain in bone, the more bone formation is
increased.

Fig : 6.1

BIOMECHANICAL RESPONSE 61
CHANGES IN CONCENTRATION OF INTRACELLULAR MEDIATORS :

Fluctuation in the concentrations of intracellular second messenger molecules has


been reported by numerous investigators. In general, cell surface receptors relay
information by activating a chain of events that alters the concentration of one or more
small intracellular signaling molecules often referred to as second messengers or
intracellular mediators. In turn, these messenger molecules pass the signal on by altering
the behavior of selected cellular proteins. Some of the most widely used intracellular
mediators arc cyclic AMP (cAMP), Ca +2, and cyclic GMP (cGMP). Prostaglandin E 2 and
Prostacyclin are paracrines that are released by osteoblasts in response to mechanical strain.

Rodan et al. (1975) agreed that the second messenger cyclic AMP was affected by
mechanical strain and also reported changes in cyclic GMP and calcium ions. Yeh and
Rodan also suggested that prostaglandins might be involved in the transduction of
mechanical strain, but did not apply physiologic levels of strain to their samples.
Osteoblasts form bone by secreting many extracellular matrix proteins, including
type I collagen, osteopontin, osteocalcin, osteonectin, bioglycan and decorin. Many of these
proteins are currently being investigated with regard to their role in the transduction of
mechanical strain.
Expression of osteopontin messenger UNA (mRNA) has been shown to be elevated
as sequelae of mechanical stress. Osteocalcin, also known as bone Gla protein, is widely
used as a marker for bone metabolism. Studies have shown that the production of
osteocalcin can be stimulated by mechanical stress both in vivo and in vitro.

CHANGES IN CELLULAR PROLIFERATION :

As previously discussed, the response of osteoblasts-like cells to mechanical strain


has been shown to be variable. Many studies have reported increases in cell proliferation,
total protein production, and DNA synthesis in response to mechanical strain.

Rodan et al (1975) also suggested that at high magnitudes of strain, osteoblasts


proliferate and there is decreased production of osteoblasts phenotypic markers, such as
alkaline phosphatase and bone matrix proteins. At lower magnitudes of strain, osteoblasts

BIOMECHANICAL RESPONSE 62
exhibit a more differentiated state, with an increase in alkaline phosphatase and matrix
protein production and but there is a decrease in proliferation.

CHANGES IN CELLULAR MORPHOLOGY AND/OR ORGANIZATION :

Ives et al. (1984) using human and bovine endothelial cells found that the cells
responded differently to various types of strain. The cells oriented themselves parallel to
the direction of shear strain induced by fluid flow, but perpendicular to the axis of
mechanical deformation on a cyclically stretched polyurethane membrane.

Investigations by Buckley et al. (1990) using osteoblast-like cells stimulated by


cyclic mechanical strain also resulted in the alignment of the cells perpendicular to the
strain vector. This perpendicular alignment was noted at 4 hours after loading and was
significant by 12 hours. A second hypothesis suggested that the cells may have resolved
their focal contacts and migrated in an attempt to minimize the strain to which they were
subjected.

ALTERED EXPRESSION AND/OR RE-ORGANIZATION OF OSTEOBLAST


INTEGRINS :

Although changes in the distribution of the cytoskeleton in mechanically strained


cells have been reported, the exact mechanism for the initial detection and transduction of
mechanical force into a biologic signal has yet to be determined. One possible transduction
pathway is the extracellular matrix—integrin—cytoskeleton axis.

Integrins are the primary receptors used by animal cells to attach to the extracellular
matrix. They function as transmembrane linkers that mediate bidirectional interactions
between me extracellular matrix and the actin cytoskeleton. Integrins are composed of two
noncovalently associated transmembrane glycoprotein subunits called alpha and beta, both
of which contribute to the binding of the matrix protein. Following the binding of a typical
Integrins to its ligand in the matrix, the cytoplasmic tail of the (3 chain binds to both talin
and a-actinin and there by initiates the assembly of a complex of intracellular attachment
proteins that link the integrin to actin filaments in the cell cortex interactions that integrins

BIOMECHANICAL RESPONSE 63
mediate between the extracellular matrix and the cytoskeleton play an important part in
regulating the shape, orientation, and movement of the cells.

Numerous subunits have been characterized and different combinations of alpha and
beta subunits function as receptors for a variety of extracellular proteins. The beta1 integrin
subunit is often expressed in bone cells both in vitro and in vivo.

Carvalho et al. (2014) demonstrated that changes in the organization of the beta1
subunit were induced by the application of strain as early as 4 hours from its onset. They
compared the expression of the β1 integrin subunit mRNA from strained cultures with
unstrained controls.

CHANGES IN GENE EXPRESSION :

To characterize the biologic response of osteoblasts like cells to external mechanical


loading, many researchers are investigating strain-induced alterations in patterns of
osteoblast gene expression. Several authors have reported that the initial response to strain
is a rapid increase in c-fos mRNA expression, indicative of increased proliferation, paired
with a rapid decline in levels of mRNA encoding bone matrix proteins, such as type I
collagen, osteopontin, and osteocalcin. The term matrix proteins refers to both collagenous
and noncollagenous proteins. Type I collagen is the most abundant protein in the organic
matrix of bone. This molecule is made up of one α2 and two α1 chains. These three chains
are initially assembled into a triple helical structure within the cell and are subsequently
bundled into fibrils once secreted from the cell. These extracellular fibrils are arranged in
a specific, repeating orientation that produces the typical banded appearance common to
type I collagen. Active osteogenesis involves the expression of genes that result in the
production of collagen type I protein. This trait makes the type I collagen molecule a
valuable indicator of differentiated osteoblastic activity.

In the last 20 years, non-collagenous proteins have received increased attention. It


has been suggested that these minor components of organic bone matrix may play a role in
regulating bone function, expression, and turnover.

BIOMECHANICAL RESPONSE 64
Osteocalcin (bone Gla protein) is a non-collagenous protein that binds calcium and
has been isolated from bone, dentin, and other mineralized tissues. It is specifically
synthesized by differentiated osteoblasts, and like type I collagen, is an ideal marker for
osteoblast phenotypic expression.

Another non-collagenous protein that is generating great interest is osteopontin.


This bone sialoprotein is synthesized by primary osteoblasts and has been shown to play a
role in cell attachment and spreading. Osteopontin contains a binding sequence that appears
to be recognized by an integrin cell surface receptor related to the vitronectin receptor.

Both osteocalcin and osteopontin are regulated by a number of hormones and growth
factors. The most common promoter of osteocalcin and osteopontin expression and
secretion is, 25-(OH)2D3, which directly effects the genes of both proteins. This is possible
because the genes for both osteopontin and osteocalcin contain regions that recognize
vitamin D.

In the absence of vitamin D, osteopontin message levels were increased several-fold


by the application of mechanical load. This increase in osteopontin expression was doubled
when the cells were subjected to mechanical load in the presence of vitamin D.

Osteocalcin secretion was also increased with cyclic strain. Osteocalcin levels were
not detectable in vitamin D untreated control cells.

Frost H (1982) has reported that the mechanisms for the biomechanical response of
osteoblasts are not discrete. Osteoblastic products such as interleukin-l (IL-I) can stimulate
osteoblasts. He groups these cells as basic multicellular units (BMUs). These BMUs are
most prevalent on periosteal and endosteal surfaces, and the periosteal BMUs are most
sensitive to biomechanical stimuli.

THE BIOMECHANICAL RESPONSE :

The mechanical properties of the trabecular and cortical bone found within the oral
environment exhibit a high degree of variation as a function of load direction, rate, and
duration. In addition, the structural density of the bone has a significant influence on its

BIOMECHANICAL RESPONSE 65
stiffness (modulus of elasticity) and ultimate strength as such, the mechanical strain
exhibited in bone is ultimately a function of the bone density.

- Dependence on Direction of Loading :

A material is said to be orthotropic if it exhibits different properties in all three


directions and isotropic if the properties are the same in all three directions. Transversely
isotropic describes a material in which two of the three -directions exhibit the same
mechanical properties.

Bone has been reported to be transversely isotropic by Reilly and Burstein (1975)
and by Yoon and Katz (1984) but Knets and Malmeisters Ashman el al. (1974) have
described bone as orthotropic (i.e., E1=E2:=E3). The mandible has been reported as
transversely isotropic with the stiffest direction oriented around the arch of the mandible.
These authors suggest that cortical bone of the mandible functions as a long bone that has
been molded into a curved beam geometry. The stiffest direction (around the arch) t hus
corresponds to the long axis of the tibia or femur. Such data raise interesting questions
regarding the primary loads that the mandible experiences: which are the occlusal loads or
flexural loads imposed during opening and closing of the mouth.

Clinical studies has qualitatively revealed that the actual mandible has a more
compact bone at the inferior border, less compact bone on the superior aspect, and greater
quality of trabecular bone, especially between the mental foramens. In addition, the
presence of teeth and/or implants significantly increases the trabecular bone amount and
density within the residual alveolar bone.

- Dependence on Rate of Loading :

A material is said to be viscoelastic if its mechanical behavior is dependent on the


rate of load application. The strain rate dependence of bone was investigated by Mc
Elhaney (1976) and reported that a significant difference can be noted in both ultimate
tensile strength and modulus of elasticity over a wide range of strain rates, with bone acting

BIOMECHANICAL RESPONSE 66
both stiffer and stronger at higher strain rates. It was restated, bone fails at a higher load,
but with less allowable elongation (deformation) at higher as compared with lower strain
rates. Thus bone behaves in a more brittle fashion at higher strain rates.

Carter and Halyes (1989) have reported both strength and elastic modulus of human
bone to be proportional to strain rate raised to the 0.06 power. Strain rate to which bone is
normally exposed varies from 0.001 sec-1 for slow walking to 0.01 sec-1 for higher levels
of activity.

- Dependence on Duration of Loading :

Carter and Caler (1989) have described bone damage or fracture caused by mechanical
stress as the sum of both the damage caused by creep or time-dependent loading and cyclic
or fatigue loading and the relative interaction of these two types of damage.

They have also reported that the creep-fracture curve for adult human bone at a
constant stress of 60 MPa . Such data raise the question of whether resorption and/or failure
in the dental bruxism or "clencher" patient may be partially (or wholly) the result of an
accumulation of creep damage.

Fatigue failure has been reported for in vivo bone by the same and associates and
by others at relatively low cycles (10 4 to 108 cycles).

- Dependence on Species and Anatomic Location :

Large variations have been noted in experimental measurements of elastic modulus


and ultimate compressive strength of trabecular bone. The trabecular bone is thus the
primary structure to dissipate and transfer loads. In the edentulous mandible, trabecular
bone is continuous with the inner surface of the cortical shell. In the dentate mandible,
trabecular bone is surrounded by a thick cortical shell and dense alveolar bone under the
teeth. Finite element models of the human mandible have shown that cortical bone plays a
major role in the dissipation of occlusal loads. Thus load patterns on trabecular bone and

BIOMECHANICAL RESPONSE 67
microstructure of trabecular bone may contribute to differences in the mechanical
behaviour of the mandible as compared with other anatomic regions.

Mechanical loads in the mandible are different from those typically experienced by
long bones. In the long bones, such as the femur and tibia, loads are primarily axial. In
contrast, muscle loads in the mandible may be large and include dorsoventral shear,
twisting about the long axis of the mandible, and transverse, increasing in magnitude from
posterior to anterior in the mandible. The regional differences observed in the mechanical
properties within the human mandible likely reflect the difference in load carried by the
different regions of the mandible With muscle attachments located posteriorly on the
mandible, the anterior mandible experiences a large moment load, and is even present in
the absence of occlusal loads, caused by the buccolingual flexure of the mandible. Thus
significantly higher densities are to be expected in the anterior as compared with posterior
mandible.

Though two- to three fold higher bite (occlusal) forces are present in the posterior
as compared with anterior mandible, both apparent density and ultimate compressive
strength of trabecular bone are lowest in the posterior mandible. These data suggest that
the large, multiple-root structure of molar teeth serves to dissipate such posterior occlusal
loads as opposed to concomitantly higher ultimate strengths in the bone itself.

Current clinical practice routinely places the same size dental implant diameter and
geometry in the posterior and anterior mandible.

- Dependence on Structural Density :

Trabecular bone is a porous, structurally anisotropic, inhomogeneous material. It


had been reported that the mechanical properties like that of, the elastic modulus and
ultimate compressive strength, exhibits upto 47% to 68% higher mean values in the anterior
(region 1) compared with the posterior region of the mandible. No differences were
observed in elastic modulus and ultimate compressive strength in the region between the
premolars and molars. Based on clinical experience with varying densities of available
trabecular bone, Misch defined two types of trabecular bone in his clinical classification
scheme for the mandible and maxilla: "coarse" (D2) in the anterior mandible and "fine"
trabecular bone in the posterior mandible (D3).

BIOMECHANICAL RESPONSE 68
Qu et al. (1986) found that there was a significant difference between apparent
density in region 1 (anterior mandible) and in regions 2 and 3 (posterior mandible) but no
significant difference was noted between region 2 and region3.

INTERFACIAL MICROMOTION AND IN VIVO TISSUE RESPONSE:

Micromotion (relative motion)—a relative displacement between an implant and


surrounding tissue—the following conclusions emerge.

First, it is not the absence of loading per se that is critical for osseointegration around
implants, but rather the absence of excessive micromotion at the interface. In this statement,
the term osseointegration might best be understood to mean simply “undisturbed bone healing
around the implant.”

Second, with micromotion, it is important to specify when the excessive micromotion


occurs relative to the time of implantation.

Micromotion, if excessive, is thought to damage the tissue and vascular structures that
are part of the early stages of bone healing. Micromotion probably interferes with development
of an adequate early scaffold from the fibrin clot. Also, micromotion probably disrupts
angiogenesis and the establishment of a new vasculature for the healing tissue, which in turn
interferes with the arrival of regenerative cells.

Third, it is especially relevant that these findings about micromotion are true not only
for metallic biomaterials (with either smooth or rough surfaces), but also for ceramic
biomaterials such as Hydroxyapatite.

The threshold at about 100 μm, assuming micromotion is started and maintained at this
magnitude soon after surgery. However, maintaining the same amount of micromotion
throughout an animal experiment can be problematic and depends on the setup.

Factors that will most likely affect stability include the shape of the implant relative to
its bone site, the surface texture of the implant, the properties of the bone, the nature of the

BIOMECHANICAL RESPONSE 69
loading on the implant, and the splinting design for the implants (if used in a full-arch situation),
among other factors

It is detrimental to have excessive micromotion at an interface, it makes sense to direct


design efforts toward developing as strong bond possible between an implant and tissue. For
example, in the extreme case of an infinite- strength interfacial bond, it would be impossible to
have any interfacial micromotion because the bone and implant would remain bonded together
at the interface, that an infinite-strength bond does not occur in reality. Many studies have
centred on the exact nature and strength of the attachment that does exist, but some papers have
questioned whether in fact this attachment has any appreciable physicochemical strength.

Fig : 6.2 (EFFECT OF MICROMOTION)

CEMENT LINE AT THE BONE-IMPLANT INTERFACE:

The biochemical and structural similarities exist between bone-biomaterial interfaces


and natural interfaces in bone itself, for example the cement line between a secondary osteon
and pre-existing bone. Davies draws attention to this sort of cement line, but more generally
for mineralized tissue, cement lines (reversal, resting lines) also exist and are actually about 10
times thinner than cement lines around osteons and or respond to what is seen around implants.

BIOMECHANICAL RESPONSE 70
Fig : 6.3 (SEM OF BONE IMPLANT INTERFACE)

CEMENT LINES IN NORMAL BONE:

Even for normal bone there is uncertainty about the exact structure and properties of
cement lines. While the osteonal cement line is known to be collagen-deficient and probably
hypo mineralized, the exact percentage of collagen, mucopolysaccharides, glycoproteins, and
mineral is still under study.

The cement line represents a residue of the ground substance in osteoid, which is
produced as a part of bone remodelling. Because the cement line represents the remnant of the
reversal phase of bone remodelling the similarities in composition [with osteoid] should not be
surprising. Significantly, cement lines in normal bone are generally thought of as weak points
in the overall composite structure. For instance, yield and fracture testing show failures at
cement lines.

Moreover, in fatigue studies of small beams of trabecular bone versus small beams cut
from dense cortical bone, Choi and Goldstein explained the poorer fatigue behaviour of

BIOMECHANICAL RESPONSE 71
trabecular bone in terms of its “mosaic” microstructure, consisting of packets of remodelled
bone separated from pre-existing bone by cement lines. However, except for these inferences
that natural cement lines are weak, little data are available on the direct mechanical properties
of cement lines.

THE MEANING OF INTERFACIAL STRENGTH DATA IN BIOMECHANICAL


MODELS OF IMPLANTS IN BONE:

Load transmission and resultant stress distribution are significant in determining the
success or failure of an implant. Factors that influence the load transfer at the bone–implant
interface include the type of loading, implant and prosthesis material properties, implant length
and diameter, implant shape, structure of the implant surface, nature of the bone–implant
interface, and quality and quantity of the surrounding bone. Of these biomechanical factors,
implant length, diameter, and shape can be changed easily. Cortical and cancellous bone quality
and quantity need to be assessed clinically and should influence implant selection.

The biomechanical relevance of data about intrinsic interface strength can be illustrated
in an Finite element study that allowed for interfacial “bond” failure according to input data
from the literature about the shear and tensile strengths of bone-implant interfaces.

The model used a typical value of 1 MPa for both the shear and tensile strengths of the
bone-titanium interface, which was assumed to consist of a cement line of essentially zero
thickness in the computer simulation. The interfacial strength values were used in a failure
algorithm for the interface.

As the implant was loaded incrementally from 0 to 300 N, stresses at the interface were
computed and compared with the defined interface failure function involving the 1 MPa limit.
The model revealed when and where the interface started to fail by debonding between implant
and bone. For the test case considered, cracks started to open early in the loading, at just 30 N,
forming at the thread cusps near the apical portion of the implant.

As loading continued up to 300 N, the cracks widened and also started at new places
along the interface, until finally at 300 N, the interface consisted of some regions remaining in
(compressive) contact because of the interlocking geometry of the threads, and other regions
where gaps had opened between the implant and bone after the interfacial tensile and shear
strength had been exceeded.

BIOMECHANICAL RESPONSE 72
Therefore, it should be recognized that the presence or absence of interfacial bonding
influences interfacial stress and strain conditions—and potentially any bone cell reactions that
might be linked somehow to the stresses and strains.

SIGNIFICANCE OF STRESS AND STRAIN IN INTERFACIAL BONE:

EXCESSIVELY HIGH STRAINS:

Typical clinical symptoms of overload include repeated prosthetic screw loosening or


fracture and loss of crestal bone.

The likelihood of a single-cycle overload failure of a bone-implant interface is not


further considered as a source of the clinically observed overload, although it is one potential
type of biomechanical overload, especially for implants in poor quality cancellous bone.

There is evidence supporting the hypothesis that fatigue microdamage can occur in
interfacial bone around a heavily loaded dental implant, and that this microdamage triggers
bone remodelling (and possibly also modelling) that may not be able to keep pace with
accumulating damage as loading continues—a situation that predisposes the bone to additional
fatigue damage and, eventually, a net loss of bone and implant failure.

DAMAGE TO BONE:

Monotonic compressive and tensile tests of both cortical and trabecular bone in the
laboratory (using regular specimens) have revealed that bone can sustain various forms of
mechanical damage when strains approach the yield point.

For cortical bone, macroscopic evidence of yielding occurs at a strain of about 0.75%,
although there is other evidence, e.g., by acoustic emission, of yielding at lower strains such as
0.5%.

For trabecular bone, the yield strain is more difficult to pinpoint because the nominal
strain of an entire specimen can differ from the strain fields that develop in individual
trabeculae, as illustrated in computer models of trabecular bone.

For bone, it is known that cracks, delamination’s, shear bands, and other phenomena
yet to be clarified comprise the nature of the microdamage seen in the microscope.

BIOMECHANICAL RESPONSE 73
In other words, it is possible that damage might be occurring in bone even though the routine
strain analyses had estimated and calculated that the nominal strain values were as safe.

DAMAGE, BONE REMODELLING, AND THE POSSIBLE RELATIONSHIP TO


‘OVERLOAD’ :

Mori and Burr (1998) established that microdamage to bone stimulates repair by bone
remodelling. Their study also indicated that :

(1) Microdamage can contribute to increased bone fragility and fracture risk.

(2) Fractures can develop as the result of a vicious cycle (positive feedback mechanism)
involving damage, remodelling-induced porosity, weakening of bone, further damage, and so
on. It is hypothesized that positive feedback occurs when bone remodelling tries to repair a
damaged site in bone, but in so doing, causes increased porosity and a vicious cycle of
worsened strain state, more damage, more remodelling, more porosity, and so on, until failure.

Loss of osseointegration was attributed to “fatigue micro fractures in the bone


exceeding the repair potential.” In an analysis of a clinical case in which overload was
suspected, Prabhu and Brunski ( 1999 ) used finite element analyses to evaluate a case in
which Brånemark implant supported prosthesis in the mandibular molar region of a human
with a positive history of bruxing. The implants had healed for 6 months before loading. After
about 2 to 3 months of function, crestal bone loss was noted radio graphically around the mesial
implant, which eventually fractured. Three-dimensional Finite element analysis predicted high
strains—in excess of 1%—at the crestal region of the mesial implant’s interface, especially for
the case of medium-size bite forces (e.g., 250 N) on the mesial cantilever of the prosthesis.

SIGNIFICANCE OF STRESS AND STRAIN IN INTERFACIAL BONE: “WOLFF’S


LAW”

BIOMECHANICAL RESPONSE 74
This “law” has been translated as “Every change in the form and function of bone(s)
or of their function alone is followed by certain definite changes in their internal architecture,
and equally definite secondary alterations in their external conformation, in accordance with
mathematical laws”- WOLFF’S LAW.

Fig : 6.4

Crestal bone loss is possible around dental implants because of abnormally low strains
and stress shielding and woven bone existed at the interfaces. The bone-implant interface is a
highly complicated milieu, in which it would be difficult to separate out the events related to
intrinsic healing from events ascribed to ‘stress shielding’.

Indeed, bone around a freshly placed implant is traumatized by implant surgery and
will be in an active state of healing for many months after surgery.

Therefore, for bone around oral and maxillofacial implants, the thought is shifting from
a blind acceptance of Wolff’s Law and moving towards a new, more testable, specific
hypothesis. For example, may be bone around oral implants is not sensitive to stress and strain
except when it is healing.

BIOMECHANICAL RESPONSE 75
Fig : 6.5

IMPLANT-PROSTHESIS INTERFACE:

The screw loosening problem is of concern, especially when considering single-tooth


implant prostheses. The application of optimal preload has been the main means of preventing
loosening. Studies advocate the addition of a washer as a simple and effective solution for the
persistent problem of screw loosening. Stress concentrations in the fastening screws are
influenced by load magnitude and direction. High rigidity prostheses and rigid abutments have
been found to give more favourable stress distributions in the screws.

Fig : 6.6

IMPLANT - NATURAL TEETH FOR FIXED PROSTHESIS:

BIOMECHANICAL RESPONSE 76
Combining natural teeth and implants to support fixed prostheses has been advocated
by certain investigators of implant dentistry. Controversy exists as to the advisability of this
design philosophy from a biomechanical as well as a clinical perspective. A significant clinical
consideration in the restoration of partial edentulism with implant- and tooth-supported
prostheses is whether implants and natural teeth abutments should be splinted, and if so, in
what manner. There is a differential deflection between the viscoelastic intrusion of a natural
tooth in its periodontal ligament and the almost negligible elastic deformation of an
osseointegrated implant. This difference may induce a fulcrum-like effect and possibly
overstress the implant or surrounding bone. Some factors that biomechanically influence the
stress distribution include abutment design, implant material properties, the effect of resilient
elements, connector design (precision or semi precision attachments), and the degree of
splinting implants to natural tooth abutments. For the implant connected with a natural tooth,
Van Rossen et al (1990) concluded that a more uniform stress was obtained around implants
with stress absorbing elements of low elastic modulus. They also concluded that the bone
surrounding the natural tooth showed a decrease in peak stresses in such a situation.

A finite elemental analysis comparing models representing a natural tooth and an


integrated implant connected by rigid and nonrigid connectors was done, on the basis of the
similarities in stress contour patterns and the stress values generated in both models, the authors
concluded that it may be erroneous to advocate a nonrigid connection because of a
biomechanical advantage.

Melo et al (2013) also investigated tooth-and implant-supported prostheses in free-end


partially edentulous situations. Their 2- dimensional Finite Element analysis predicted that
lowest levels of stress in bone occurred when the prosthesis was not connected to a natural
abutment tooth but instead was supported by two freestanding implant abutments. Nonrigid
attachments, when incorporated into a prosthesis, did not significantly reduce the level of stress
in bone. A recent comprehensive review of both clinical and laboratory studies concluded that
the issue of connecting natural teeth to implants with rigid or nonrigid connectors remains
unresolved.

BIOMECHANICAL RESPONSE 77
FORCE DELIVERY AND FAILURE MECHANISMS

A consensus stated that, the location and magnitude of occlusal forces affect the quality

and quantity of induced strains and stresses in all components of the bone-implant prosthesis

complex. When evaluating the biological effects of an applied load, it is essential to determine

its source. Qualification and quantification of these forces on implants and in bone is required

to understand the in vivo behaviour of these devices. So far, in vivo studies have measured

forces on implants have been measured only at the abutment level. Since intraosseous strains

in the vicinity of implants have not been measured by means of biosensors, strain gradients that

guide bone modelling and remodelling processes around implants are unknown. Because

correct evaluation of forces is often a perplexing problem and a challenge to resolve due to

several accompanying parameters involved in experiments, correct in vivo isolation of forces

in the vicinity of implants are always avoided. As a result, obtaining an undisputed scientific

proof becomes impossible.

The manner in which forces are applied to implant restorations within the oral

environment dictates the likelihood of system failure. The duration of a force may affect the

ultimate outcome of an implant system. Relatively low-magnitude forces, applied repetitively

over a long time, may result in fatigue failure of an implant or prosthesis. Stress concentrations

and, ultimately, failure may develop if insufficient cross-sectional area is present to dissipate

high-magnitude forces adequately. If a force is applied some distance away from a weak link

in an implant or prosthesis, then bending or torsional failure may result from moment loads.

An understanding of force delivery and failure mechanisms is critically important to the

implant practitioner to avoid costly and painful complications.

FORCE DELIVERY AND FAILURE MECHANISMS 78


Fig : 7.1 (FORCE COUPLE ACTING ON CANTILEVER)

MOMENT LOADS

Fig : 7.2

o In the above figure the moment of a force (m) is defined as a vector, the magnitude of

which equals the product of the force magnitude multiplied by the perpendicular

distribution (moment arm) from the patient of interest to the line of action of the force.

o Also torque / torsional load is destructive to implant system, they result in interface

breakdown, bone resorption, screw loosening, bar/bridge fracture.

FORCE DELIVERY AND FAILURE MECHANISMS 79


In all incidences of clinical loading, occlusal forces are first introduced to the prosthesis
and then reach the bone implant interface via the implant. So far, many researchers have,
therefore, focused on each of these steps of force transfer to gain insight into the
biomechanical effect of several factors such as:

• Force directions.
• Force magnitudes.
• Prosthesis type.
• Prosthesis material.
• Implant design.
• Number and distribution of supporting implants.
• Bone density.
• The mechanical properties of the bone-implant interface.

• Clinical moment arms:


As already described earlier a total of six moments may develop about the three clinical
co-ordinate axes.

Mesiodistal axis – Lingual / Facial movement

Faciolingual axis – Occlusal / Apical movement

Vertical Axis – Lingual Transverse / Facial Transverse movement.

So, we have three clinical moment arms -

1) Occlusal height.
2) Cantilever length.
3) Occlusal width.

FORCE DELIVERY AND FAILURE MECHANISMS 80


1) OCCLUSAL HEIGHT:

The occlusal height serves as the moment arm for force components directed along the
faciolingual axis (A), working or balancing occlusal contacts, tongue thrusts, or in
passive loading by cheek and oral musculature (B), as well as force components
directed along the mesiodistal axis.

Treatment planning must take into account this initially compromised biomechanical
environment. The moment contribution of a force component directed along the vertical
axis is not affected by the occlusal height because no effective moment arm exists.
Offset occlusal contacts or lateral loads, however, introduce significant moment arms.

Fig : 7.3

FORCE DELIVERY AND FAILURE MECHANISMS 81


2) CANTILEVER LENGTH:

Fig: 7.4

- Large moments may develop from vertical axis force components with cantilever

extensions or offset loads from rigidly fixed implants.

- A lingual force component also may, exists.

- Force applied directly over the implant does not induce a moment load or torque.

- When a full-implant supported prosthesis with cantilever segment supported by anterior

four or six implants, infinite number of loading cycles can be maintained at low stress

levels.

Geometry:

- It influences the degree to which it can resists bonding and torsional loads and

ultimately fatigue fracture.

- Implants rarely display fatigue fracture under axial compressive loads compared to

lateral loads.

- It also includes the thickness of metal or implant.

FORCE DELIVERY AND FAILURE MECHANISMS 82


o Fatigue fracture is related to fourth power of the thickness difference.

o Often a weak link in an implant body design is affected by the difference in the

inner and outer diameter of the screw and the abutment screw space in the

implant.

Force magnitude:
- There should be reduction of applied load (stress) as low as possible.

- If an applied load (stress) can be reduced, the likelihood of fatigue is reduced. As

described previously, magnitude can be reduced by.

o Higher loads on posterior compared to anterior segments.

o Elimination of moment loads.

o Optimize geometry for functional area.

o Increase the number of implants used.

Loading cycles:
- Fatigue failure is reduced to the extent if the number of loading cycles can be reduced.

Aggressive strategies to eliminate parafunctional and reduced occlusal contacts should be


taken in considerations in some cases to protect against fatigue failure.

3) OCCLUSAL WIDTH
Wide occlusal tables increase the moment arm for any undesirable offset occlusal loads.
Faciolingual tipping (rotation) can be reduced significantly by narrowing the occlusal tables or
adjusting the occlusion to provide more centric contacts.

In summary, a vicious, destructive cycle can develop with moment loads and result in
crestal bone loss. As when crestal bone loss develops, occlusal height moment arm

FORCE DELIVERY AND FAILURE MECHANISMS 83


automatically increases. With an increased occlusal height moment arm, the faciolingual
microrotation and rocking increase and cause even more stress to the crestal bone. Unless if
the bone increases in density and strength, the cycle continues to spiral toward implant failure
if the biomechanical environment is not corrected.

FATIGUE FAILURE

Implants are load-bearing devices aimed to function under the complex nature of

mastication loads. Prostheses supported by dental implants are subjected to various forces, and

moments, all transmitted to the implant body and its components. The force applied to an

implant is extremely variable, and its magnitude depends on the patients’ characteristics like

that of (age, gender, oral habits, etc.), type of prosthesis (single crown, overdenture, fixed

partial denture (FPD), cantilever, number and position of the implants, as well as type of food

consumed (carrots, meat etc.). The magnitude of the reported bite/mastication loads ranges

everywhere from 100 to 2400 N.

Fatigue failure is characterized by dynamic, cyclic loading conditions. Four fatigue factors

significantly influence the likelihood of fatigue failure in implant dentistry:

(1) Biomaterial.

(2) Microgeometry.

(3) Force Magnitude.

(4) Number of Cycles.

Fatigue behaviour of two biomaterials is characterized graphically in what is referred

to as an S-N curve (a plot of applied stress versus number of loading cycles) in Fig: 7.5 . If an

implant is subjected to an extremely high stress, then only a few cycles of loading can be

FORCE DELIVERY AND FAILURE MECHANISMS 84


tolerated before fracture occurs. Alternatively, an infinite number of loading cycles can be

maintained at low stress levels. The stress level below which an implant biomaterial can be

loaded indefinitely is referred to as its endurance limit. Titanium alloy (A) exhibits a higher

endurance limit than Commercially Pure Titanium (B).

The geometry of an implant also influences the degree to which it can resist bending

and torsional loads and ultimately fatigue fracture. Implants rarely, may display fatigue fracture

under axial compressive loads. Morgan et al (1999) reported fatigue fractures of Brånemark

dental implants caused by cyclic buccolingual loads (lateral loading) in an area of weak

bending strength within the fixture (i.e., reduced moment of inertia). The fracture of the

implant body occurred in three of the patients studied, and fracture of the abutment screws for

the Brånemark implant occurred in fewer than three patients. Fifteen acrylic or composite tooth

fractures occurred on 10 to 20 of the fixed prostheses supported by implants over a 1- to 5-year

period.

The geometry also includes the thickness of the metal or implant. The fatigue fracture

is related to the fourth power of the thickness difference. A material two times thicker in wall

thickness is approximately 16 times stronger. Even small changes in thickness can result in

significant differences. Often the weak link in an implant body design is affected by the dif-

ference in the inner and outer diameter of the screw and the abutment screw space in the

implant.

To the extent that an applied load (stress) can be reduced, the likelihood of fatigue

failure is reduced. As described previously, the magnitude of loads on dental implants can be

reduced by careful consideration of arch position (i.e., higher loads in the posterior compared

with anterior mandible and maxilla), elimination of moment loads, and increase in surface area

FORCE DELIVERY AND FAILURE MECHANISMS 85


available to resist an applied load (i.e., optimize geometry for functional area, or increase the

number of implants used).

Fig : 7.5

FORCE DELIVERY AND FAILURE MECHANISMS 86


OVERLOADING RISK FACTORS FOR DENTAL
IMPLANTS

1) Parafunction:

Parafunctional forces on teeth or implants are characterized by repeated or sustained,


occlusion and have long been recognized as harmful to the stomatognathic system. The
most common cause of early loss of rigid fixation during the first year of implant loading
is the result of parafunction. Such complications occur with greater frequency in maxilla,
because of a decrease in bone density and an increase in the moment of force. The lack of
rigid fixation during healing is also often a result of parafunction on soft tissue borne
prostheses overlying the submerged implant. Hence the presence of these conditions must
be carefully noted in implant dentistry. Most commonly encountered parafunctions are :

• Clenching.
• Bruxism.
• Tongue thrusting.

Fig: 8.1 Fig : 8.2

2) Disequilibrium Masticatory Dynamics:

Masticatory muscle dynamics are responsible for the amount of force exerted on the
implant system. The force is related to the amount and duration of function. For example,
chewing paraffin wax for 1 hour each day for 1 month can increase the biting force in men
from 118 to 140 lb. Chewing gum frequently can cause a similar increase. Eskimos, whose
diets include extremely tough substances, reached values of about 300 psi maximum forces.

OVERLOADING RISK FACTORS 87


Parafunctional bruxism or clenching often leads to hypertrophy of the muscles of
mastication and increased force.

The size of the patient can influence the amount of bite force. Large athletic men can
generate greater forces; patients of weak physical condition often develop less force than
athletic patients. In general, the forces recorded in women are 20 lb less than those in men.
Older patients record lower bite force than young adults. In addition, the younger patient
needs the additional implant support for the prosthesis for a longer time. An 80-year-old
patient will need implant support for far fewer years than a 20-year-old, with all other
factors equal.

The maximum bite force decreases as muscle atrophy progresses throughout years of
edentulousness. A maximum force of 5 psi may be the result of 30 years edentulousness.
This force may increase 300% in the 3 years following implant placement. Therefore sex,
muscle mass, exercise, diet, state of the dentition, physical status, and age may all influence
muscle strength, masticatory dynamics, and there- fore maximum bite force.

3) Opposing Arch:

Natural teeth transmit greater impact forces through occlusal contacts than do
soft tissue-borne complete dentures. In addition, the maximum occlusal force of
patients with complete dentures is reduced and may range from 5 to 26 psi. The force
is usually greater in recent denture wearers, and decreases with time. Muscle atrophy,
thinning of the oral tissues with age or disease, and bone atrophy often occurs in the
edentulous patient as a function of time. Some denture wearers may clench on their
prosthesis constantly, which may maintain muscle mass. However, this condition
usually accelerates bone loss. Implant overdentures improve the masticatory
performance and permit a more consistent return to centric relation occlusion during
function. The maximum force is related to the amount of tooth or implant support.
Partial denture patients may record forces of approximately 26 psi, which is
intermediate between that of natural teeth and complete dentures and depends on the
location and condition of the remaining teeth, muscles, and joints. In the partially

OVERLOADING RISK FACTORS 88


edentulous patient with implant-supported fixed prostheses, force ranges are more
similar to those of natural dentition.

Fig : 8.3

4) Stress Multipliers:

Stress multipliers magnify the total force generated in a system as a result of a


lever action or torque. Failures of natural teeth as a result of excessive abnormal stress
are attributed primarily to leverage and torque rather than to individual tooth occlusal
overload.

5) Direction of Load:

The direction of the occlusal load results in significant differences in the amount
of force exerted on an implant. Forces are tensile, compressive or shear to the implant
system. Three-dimensional stress analysis has shown that almost all the stresses occur
in the coronal half of the implant and bone. There is much less tensile and compressive
stress with vertical loads compared to an angled load on an implant. A lateral load on
an implant crown makes the crown height act as a lever and force magnifier. Lateral
forces represent approximately a 50% to 200% increase in stress compression compared
with vertical loading, and tensile stresses with horizontal stress increased more than
tenfold. Since early crestal bone loss occurs in similar fashion as these stress contours,
methods to reduce crestal stress are aimed at improving implant health and longevity.

OVERLOADING RISK FACTORS 89


The direction of forces may be one of the more critical factors to be evaluated during
implant treatment planning.
Maxillary anterior teeth or implants are rarely placed in the direction of occlusal
forces. After tooth loss, the bone is resorbed from the labial aspect first, and labial
concavities are often present in this region. Therefore the implant apex must be placed
with a palatal angulation. Mandibular posterior implants are often placed with a facial
inclination of the implant apex, to avoid perforation of the submandibular fossa. These
anatomic configurations affect implant angulation and the final treatment plan. If the
forces of occlusion are not axial to the implant body, additional implants, wider
implants, stress relievers in the prosthesis, or overdentures should be considered.

6) Crown Height:

The crown height also affects the amount of forces distributed to the implant-
prosthetic system in the presence of lateral or cantilevered forces. The greater the crown
height, the greater is the moment of the force under lateral loads. The crown height acts
as a lever with any lateral force. Since stresses are concentrated at the crest of a rigidly
fixated implant, the crown height multiplier increases stress rapidly. For every 1mm
crown height increase, a force increase may be 20%. Therefore, a crown height increase
from 10 to 20 mm may increase stress 200%.
The vertical distance from the occlusal plane to the opposing landmark for
implant insertion is typically a constant in an individual. Therefore, as the bone resorbs,
the crown height becomes larger, but the available bone height decreases. An indirect
relationship is found between the crown and implant height.
Moderate bone loss results in a crown-implant ratio greater than 1, and results
in greater forces being applied to the crestal bone with lateral force than in abundant
bone, in which the crown height is less. A linear relationship exists between the applied
load and internal stresses. Therefore, the greater the load applied the greater the tensile
and compressive stresses transmitted at the bone interface and to the prosthetic
components.

OVERLOADING RISK FACTORS 90


Fig : 8.4
7) Bone quality and quantity :

Lower implant survival rates have been associated with reduced quantity and
quality of bone. Hermann et al (2005) reported a 5-year survival rate of implants placed
at different configuration of bone quality and quantity. Healthy patients who presented
with a combination of type IV bone and jaw shape of Class D & E (Classified according
to Lekholm et al 1985) experienced up to 63% failure rate. In addition, Noguerol et al.
(2006) noted a 1.93 greater risk of early implant failure in implants placed in anything
but type II bone. Due to differences in bone quality between the maxilla and mandible,
location of implants may be associated with higher or lower survival rates.

Fig : 8.5

OVERLOADING RISK FACTORS 91


Implants placed in the maxilla have failure rates three times higher than those
in mandibular implants. More than half of the early failures occur in the posterior of the
maxilla. Often the maxillary posterior region lacks thick cortical plate, dense medullary
bone, and vertical dimension due to sinus pneumatization. The poor quality of bone of
the maxillary posterior has been cited as a reason for the lack of initial stability and
high early failures. This association between quantity/quality of bone and implant
survival rates emphasizes the importance of and prerequisite for bone augmentation
procedures prior to implant placement in atrophied jaws. Tilted implants, however,
have been proposed as an alternative to axially placed implants in regenerated bone.
Evaluation of survival rates of tilted implants was examined in a recent in which there
was no significant difference in the implant failure rate or the marginal bone loss.

8) Cantilever :

Excessive Cantilever : Since the introduction of the implant supported


cantilever prosthesis for the completely edentulous arch, studies have showed that the
cantilever has become a more accepted modality in implant dentistry for partially
edentulous patients Romanos G et al (2012), it places offset loads to the implant
abutments in greater tensile and shear forces on cement or screw fixation.

Fig : 8.6

OVERLOADING RISK FACTORS 92


The placement of implants is a crucial factor to consider in a three unit posterior
prosthesis. If such prosthesis is supported by two implants and has a cantilevered tooth,
the bending moment maybe twice that of a prosthesis in which both ends are supported.
With occlusal forces acting on the cantilever, the implant becomes a fulcrum and is
subjected to axial rotational, torsional forces.

Pier Abutments : There is difference in mean axial displacement between


natural teeth and dental implant because of which placing an implant in a pier situation
is significant. The breakdown of supporting tissues is extremely rapid because the
dental implant will take most of the load as a result of the differences in mean of axial
displacement. The most common solution given to overcome this is use of non-rigid
connectors. Misch stated that when an implant serves as a pier abutment the natural
tooth may become mobile over the time because the implant may act as a fulcrum. He
recommended use of the stress-breaking element as well.

Fig : 8.7

9) Non-Passive Fit :

A passive fit reduces long term stresses in the superstructure and the implant
components and also in the adjacent bone. The absence of passive fit maybe manifested
clinically by pain and discomfort in the short term, and loosening or fracture of implant
components in the long term because of the excessive strain on the adjacent bone.

OVERLOADING RISK FACTORS 93


10) Improper Occlusal Scheme :

Occlusal factors are a primary requisite for long term survival, because a poor
occlusal pattern increases and localizes forces. These factors may lead to more frequent
complications of the prostheses and bone support. The occlusal pattern of dental
implants was derived from the basic occlusal concepts of natural teeth. However, the
occlusal trauma on dental implants is more offensive than on natural teeth because of
the force dissipation difference and because of differences in proprioception.

11) Bending Movements :

Bending overload can be defined in a situation in which occlusal forces on an


implant-supported prosthesis, exert a bending movement on the implant cross section
at the crestal bone, leading to marginal bone loss and/or eventual implant fatigue.
Quirynin et al (1995) and Hoshaw et al (2015) have shown both clinically and
experimentally, that both resorption around an implant maybe caused by overload. This
will induce bending moments on the implant and three causative factors were
enumerated :
• Implants in a straight line.
• Leverage.
• Bruxism And Heavy Occlusal Forces.

fig : 8.8

OVERLOADING RISK FACTORS 94


12) Premature Loading :

Rapid loading is considered to the one of the most common causes of


prosthetic related implant failure. Branemark protocol requires a strict stress
free healing period of 3-6 months for osseointegration to occur. Brunski (1992)
stated that micromotion of more than 100 micrometers should be avoided.
Motion greater than this level would cause the wound to undergo fibrous tissue
repair rather than the desired osseous regeneration. But this is hard to apply
practically and it can be further said that the actual tolerable micromotion is
unknown.

OVERLOADING RISK FACTORS 95


SCIENTIFIC RATIONALE FOR IMPLANT &
PROSTHESIS SELECTION

Dental implants function to transfer load to surrounding biologic tissues. Thus, the

primary functional design objective is to manage (dissipate and distribute) biomechanical

loads to optimize the implant supported prosthesis function. Biomechanical load

management is dependent on two factors:

1. The character of the applied force.

2. The functional surface area over which the load is dissipated.

There are more than 50 dental implant body designs available, A scientific rationale of

dental implant design may evaluate these designs as to the efficacy of their biomechanical

load management.

Physiologic Constraints on Design:

Normal physiology limits on-the magnitude of forces that must be withstood by

engineering designs in the Oral environment. The magnitude of bite force varies as function

of anatomic region and state of the dentition. The magnitude of force is greater in the molar

region (200 lb}, less in the canine area (100 lb), and least in the anterior incisor region 25

to 35 lb). These average bite forces increase with parafunction to magnitudes that may

approach 1000 lb.

Its ultimate strength is highly dependent on its density. As such, less dense bone

may no longer be able to support normal physiologic bite forces on implants. In addition,

studies on dentate and edentulous mandibles illustrate greater trabecular bone density in

the anterior mandible, compared with the premolar or molar region.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 96


Influence on Biomaterial Selection:

Many biocompatible materials are unable to withstand the magnitude of

parafunctional loads that may be imposed on dental implants. Materials such as

silicone. Hydroxyapatite and carbon are characterized by ultimate strengths that are too low

when used as a primary implant biomaterial, even though they are quite biocompatible with

biologic tissues.

Titanium and titanium alloy have a long history of successful use in dental and

orthopedic applications. The excellent biocompatibility of titanium and its alloy has been

confirmed. With its highly active TiO2 layer, the material is extremely well tolerated by

local tissues. Titanium-aluminum-vanadium (Ti-6AI-4V) alloy has been shown to exhibit

the most attractive combination of mechanical and physical properties, corrosion

resistance, and general biocompatibility of all metallic biomaterials. The primary advantage

of titanium alloy as compared with commercially produced titanium (CpTi) is its strength.

Mechanical Properties for Commercially Produced-Titanium and Titanium Alloy

PROPERTY GRADE 1 GRADE2 GRADE 3 GRADE 4 Ti-6AI-4V

Tensile strength 240 345 450 550 930

(MPa)

Yield strength 200 170 275 380 483

(MPa)

Modulus of 103 103 103 103 113

elasticity

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 97


Titanium and its alloy represent the closest approximation to the stiffness of bone

of any surgical grade metal used as an artificial replacement for skeletal tissue, even though

it is almost 6 times more stiff than dense cortical bone. Thus, titanium alloy represents the

best compromised solution (given current biomaterials technology) between biomechanical

strength, biocompatibility, and the potential for relative motion (from modulus mismatch)

at the bone-to-implant interface.

Clinical Implant Design Failures Related to Choose of Biomaterial and Force

Magnitude:

Two examples of implant body failures related to biomaterial choice have appeared

in the historical implant literature. The vitreous carbon implants optimized the modulus of

elasticity (stiffness) of the biomaterial (carbon) without appropriate attention to ultimate

strength considerations. Conversely, Al 2O3. Ceramic implants optimized ultimate strength

without adequate attention to modulus of elasticity.

The vitreous carbon implant design was composed of a carbon body with an internal

316-L stainless steel post. The stiffness of the carbon was compatible with the surrounding

bone; however, the carbon body was incapable of withstanding the physiologic loads within

the oral environment. The post was then subjected to dramatic corrosion with the

subsequent release of metallic ions into the interfacial tissues. A close match of biomaterial

and bone material stiffness alone cannot, in isolation, provide clinical success.

The ceramic implants, as a class, were antithetical to the carbon implants. Ultimate

compressive strength was optimized at the sacrifice of matching biomaterial and bone

stiffness. The modulus of elasticity for ceramics is approximately 33 times stiffer than bone.

The very stiff ceramic implants carried a disproportionate amount of the load and the

interfacial bone was moved into disuse atrophy.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 98


Physiologic Constraints on Design:

The duration of bite forces on the dentition has a wide range. Under ideal conditions,

the teeth come together during swallowing and eating for only brief contacts. The total time

of those brief episodes is less than 30 minutes per day. Patients who exhibit bruxism,

clenching, or other parafunctional habits, however, may have their teeth in contact several

hours each day.

Influence on Implant Body Design:

Materials that are subjected to repetitive loads are at greater risk of fatigue failure.

Mechanical stress may be great enough in magnitude to fracture a material al one cycle

(i.e., one application of load). If the material receives less stress, it may st ill fracture, but

after more cycles. The endurance limit or fatigue strength is the level of highest stress a

material may be repetitively cycled without failure. The endurance limit of a material is

often less than one half its ultimate tensile strength. Hence fatigue and ultimate strength

values are related, but fatigue is a more critical factor, especially for patients with

parafunction since they impose higher stress magnitude and greater cycles of load

Off axis, cyclic loading of an implant or its prosthetic components, even with a

relatively low magnitude of force, can also cause failure and/or fracture of the implant

components. Dental implants are designed for loading along their long axis and the implant

body is particularly susceptible to fatigue fracture with bending loads in the buccolingual

plane. Such transverse bending loads may be caused by premature contacts, bruxism, or

significantly angled implants. No root form implant is specifically designed to withstand

cyclic bending loads; therefore great -caution in treatment planning must be taken to avoid

destructive transverse and/or bending loads to implants.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 99


The ability of implants and abutment screws to resist fracture from bending loads is

directly related to the component's moment of inertia (or bending fracture resistance factor).

Implant bodies are particularly susceptible to fatigue fracture at the apical extension of the

abutment screw within the implant body or at the crestal module-around abutment screw,

which does not have direct contact (e.g., with an internal hex)- The formula for the bending

fracture resistance in these conditions is related to the outer diameter radius to the fourth

power, minus the inner diameter radius to the fourth power. Even a small increase in wall

thickness can result in a significant increase in bending fracture resistance since the

dimension is multiplied to a power of four. When the outer diameter increases 0.1mm and

the inner diameter remains-unchanged, the bending fracture resistance increases to 2.967

or a 33% increase in strength. When the outer diameter remains unchanged and the inner

diameter decreases 0.1mm, the increase is 2.671 or a 20% increase. Hence an increase in

outer diameter (which also increases overall surface area of bone support) has a more

significant effect on body wall strength

A prosthesis or coping screw often has smaller moment of inertia than its mating

implant body %(R4). Thus, if the prosthesis screw is partially loose and thereby bearing a

large component of a transverse load to the occlusal surface, the screw will fracture because

of bending fatigue. Some investigators have suggested the phenomenon of screw breakage

to be a long-term advantage for the implant. Restated, it is better for the screw to break than

the implant because the screw is easily retrievable; the implant body is not. Although this

concept has some value, it is also a faulty safety factor. Most implant prostheses have more

than one implant abutment. As soon as one screw loosens or breaks, the stresses are

increased to the remaining implants, components, and bone interfaces. The additional

cantilever loads increase the stresses and may contribute to bone loss and/or implant

component fracture.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 100


A smooth cylinder implant body results in essentially a shear type of force, at the implant -

to-bone interface". Thus, this body geometry must use a microscopic retention system by

coating the implant with titanium or HA). The integrity, of the interface is therefore

dependent upon the shear strength of the HA-to-bone bond.

Threaded implants have the ability to transform the type of force imposed at the

bone interface through careful control of thread geometry. Thread shape is particularly

important in changing force type at the bone interface. Thread shapes in dental implant

designs include square, V shape, and buttress under axial loads to a dental implant, a V

thread face (typical of Paragon, 3i, and Nobel BioCare) is comparable to the buttress thread

(typical of Steri-Oss) when the face angle is similar and has approximately a 10 times

greater shear component of force than a square or power thread (typical of BioHorizons).

A reduction in shear load at the thread-to-bone interface reduces the risk of overload, which

is particularly important in compromised D 3 and D4 bone.

As the angle of load increases, the stresses around the implant increase, particularly

in the vulnerable crestal bone region. As a result, virtually all implants are designed for

placement perpendicular to the occlusal plane. This placement allows a more axial load to,

the implant body and reduces the amount of crestal stress.

Any smooth shear surface on an implant body is at risk for bone loss because of

inadequate load transfer, depicts one such example (Core-Vent/Paragon implant)

characterized by extensive crestal resorption adjacent to a long, smooth shear surface on

the implant body. This contributed to an increase in crown height (which further magnifies

stress) and the fracture of two abutments.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 101


The face angle of the thread or plateau can change the direction of load from the

prosthesis to abutment connection, to a different force direction at the bone.

Clinical Implant Design Failures Related to Force Duration:

Morgan et al. reported on fatigue failures of Branemark fixtures subjected to

bending loads. Fixture fracture occurred, as predicted, in the region of the implant that was

characterized by a reduced annular cross-section.

Physiologic Constraints on Design:

Three types forces may be imposed on dental implants within the oral environment:

compression, tension, and shear. Bone is strongest when loaded in compression, 30%

weaker when subjected to tensile forces, and 65% weaker when loaded in shear. Endosteal

root-form implants load the bone to implant interface in pure shear (e.g. smooth sid ed

cylinder) unless surface features arc incorporated in the design to transform the shear loads

to more resistant force.

The anatomy of the mandible and maxilla places significant constraints on the

ability to surgically place root form implants suitable for loading along their long axis.

Bone undercuts further constrain implant placement and thus force direction. Mostly all

undercuts occur on the facial aspects of the bone, with the exception of the submandibular

fossa in the posterior mandible. Hence implant bodies are often, angled to the lingual, to

avoid penetrating the facial undercut during insertion. Bone is strongest when loaded in its

long axis in both compression and tensile forces. A 30-degree offset load reduces the

compressive strength of bone by 11%, and reduces the tensile strength by 25%.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 102


Fig : 9.1

Buttress Square
V-shaped

Fig: 9.2

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 103


Force Magnification:

A surgical placement resulting in extreme angulation of the implant and/or a patient

exhibit parafunctional habits will likely exceed the capability of any dental implant design

to withstand physiologic loads. Cantilevers and crown heights are levers and therefore force

magnifiers. Careful treatment planning with special attention to the use of multiple implants

to increase functional surface area is indicated when a clinical case presents the challenge

of force magnifiers. A magnifier of force around an individual implant is also related to the

density of bone. Since density is directly related to bone strength, and D4 bone is estimated

more than 10 times weaker than D1 bone, the effect of this resultant force is magnified as

to its clinical result when placed upon softer bone types.

SURFACE AREA:

Anatomic Constraints on Surface Area Optimization:

The normal anatomy of mandible & maxilla imposes significant geometric

constraints on the size and configuration of dental implants.

Bone Volume (External Architecture of Bone):

The volume of available bone is dependent on anatomic location as well as the

degree of bone resorption. The original bone volume in width is greater in the posterior

regions of the mouth. As a general rule, the bone width is more often 6 mm than 8 mm in

the anterior regions of the mouth. Hence 4-mm diameter implants are the most frequently

used in this location. The posterior regions of the mouth more often have bone widths

greater than 7 mm, and as a result implants 5mm in diameter may be used. Therefore,

implant width increases as amount of force magnitude increases from anterior to posterior.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 104


To the contrary, the bone height usually decreases from the anterior mandible compared

with the anterior maxilla, posterior mandible, to the least in the edentulous posterior

maxilla. Hence as the occlusal forces increase, the bone height (and volume) decreases.

Thus, careful and innovative engineering design is required to optimize implant design for

functional loads within these anatomic limits of bone volume.

Bone Quality (Internal Architecture of Bone):

Four distinctly different bone density classifications exist within the maxilla and mandible,

with a broad range of biomechanical strengths (i.e., ability to withstand physiologic loads).

Significantly increased clinical failure rates in poor quality, porous bone compared with

more dense bone have been documented worldwide. Failure rates as high as 35% have been

reported in D4 (Type IV) bone, and are mostly caused by early implant failures, which are

caused by overload. In order to decrease stress, the practitioner may elect to increase the

number of implants or use an implant design with greater surface area.

Design Variables in Surface Area Optimization:

Implant Macrogeometry:

Smooth-sided, cylindrical implants provide ease in surgical placement; however, the bone-

to-implant interface is subjected to significantly larger shear conditions. In contrast to

smooth-sided, tapered implant allows for a component of compressive load to be delivered

to the bone-to-implant interface, dependent upon die degree of taper. The larger the taper,

the greater the component of compressive load delivered to the interface. Unfortunately,

the amount taper cannot be greater than 30 degrees or the implant body length is

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 105


significantly reduced along with the immediate fixation required for die initial healing. In

contrast, threaded (or plateaued) implants with circular cross-sections provide for ease of

surgical placement and allow for greater functional surface area optimization to transmit

compressive loads to the bone-to-implant interface. In addition, a threaded implant is easily

rigidly fixated initially to limit micromovement during healing. A smooth-sided cylinder

depends on a coating or microstructure for load transfer to bone. This surface treatment

may also be applied to a screw or plateau design increasing the functional surface from

both design and surface treatment conditions.

Implant Width:

Over the past five decades of endosteal implant history, implants have gradually

increased in width. The Branemark implants system is of of 3.75 mm. Today,dental

implants generally have reflected the scientific principle that an increase in implant width

adequately increases the area over which occlusal forces may be dissipated. A 4 -mm root

form implant has 33% greater surface area than a 3-mm root form implant. It is important

to place the largest diameter implant fixture appropriate for the ridge width. This reduces

the effective length of the cantilever, reducing the potential for off-axis loading1. This

trend is also noted in natural teeth to compensate for increased force; molar teeth arc wider

than incisors. The larger the width of the implant, the more it resembles the emergence

profile of the natural tooth. Since most teeth are 6 to 12 mm in width, a clinical desire is to

have implants of similar size. However, the titanium implant is 5 to 10 times more rigid

than a natural tooth. The increased width of implants 6 to 12 mm affects the bending

resistance of the implant related to the radius raised to the fourth power.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 106


The implants were so rigid because of their size and biomaterial that inadequate strain

was transmitted to bone, which resorbed. This condition was also observed with the

aluminium oxide dental implants, which were 33 times more rigid than bone. Likewise,

implants of similar dimension to the premolar and molars may be too rigid to strain the

bone within physiologic ranges, and disuse atrophy may ensue. Crestal bone anatomy,

however, typically constrains implant width to less than 5.5 mm, except in limited clinical

situations. It was also found that pin implant Sialom system were less than 2mm wide than

Branemark system.

Thread Geometry:

Functional surface area per unit length of the implant may be modified by varying

three thread geometry parameters; thread pitch, thread shape & thread depth

Fig : 9.3

Thread pitch, is defined as the distance measured parallel with its axis between

adjacent thread forms (for V-type threads), or the number of threads per unit length in the

same axial plane and on the same side of the axis. The smaller (or finer) the pitch, the more

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 107


threads on the implant body for a given unit length, and thus the greater surface area per

unit length of the implant body. Restated, a decrease in the distance between threads will

increase the number of threads per unit length. Therefore, if force magnitude is increased

or bone density decreases, the thread pitch may be decreased to increase the functional

surface area. The fewer the threads, the easier to bone tap and/or insert the implant. If fewer

threads are used in stronger bone, the implant ease of placement is improved, since hard

bone is more difficult to prepare for threaded implant placement.

The thread shape is another very important characteristic of overall thread

geometry. As described previously, thread shapes in dental implant designs include: square,

V-shape and buttress. In conventional engineering applications, the V-thread designs is

called a "fixture" and is primarily used for fixturing metal parts together. It helps in

optimized pull out loads.

Fig : 9.4

The thread depth refers to the distance between the major and minor diameter of the

thread. Conventional implants provide a uniform thread depth throughout the length of the

implant. The thread depth may be varied, however, over the length of the implant to provide

increased functional surface area in the regions of highest stress (e.g., the crestal region of

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 108


alveolar bone). Specifically, a reverse taper in the minor diameter of threaded implant can

produce an increased thread depth at the, top of the implant body relative to the apex. This

unconventional design feature results in dramatic increase in functional surface area at the

crest of the bone, where the stresses are highest. Thread geometry is so powerful a mediator

of load transfer, careful attention to thread design can override a perceived advantage of

wider and/or tapered implants.

Implant Length:

As the length of an implant increases, so does the overall total surface area. The

opposing cortical plate is engaged primarily in the anterior regions of the mouth, especially

the anterior mandible. Yet, the bite forces are lower and the bone density is greater in the

anterior regions. Bi-cortical stabilization, a rationale often cited for longer implants, is

simply not needed in Dl bone because it is already a homogenous cortical bone.

In poor D3 and D4 quality bone, functional surface area must be maximized to

optimally distribute occlusal loads. Conventional thinking suggests that longer implants

provide maximum functional surface area. Yet, D3 and D4 bone are primarily observed in

die posterior regions of die jaw, where less available bone is observed compared with the

anterior regions. Nerve repositioning is cited as an acceptable clinical treatment to facilitate

placement of longer implants in the posterior mandible.

In order to place the longest implants in the maxillary posterior regions a sinus graft

is often required. Hence increasing surface area primarily by length in the posterior regions

of the jaws requires advanced grafting or nerve repositioning surgery and does not benefit

the primary regions of increased stress the crestal bone region.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 109


Longer implants have been suggested to provide greater stability under lateral

loading conditions. Finite element analysis provides an analytical means to investigate the

influence of implant length relative to functional surface area under such extreme loading

conditions. The results of this analysis point to the fact that the majority of the maximum

stress generated by a lateral load can be dissipated as well by implants in the range of 10

to 15mm in length, compared with implants in the range of 20 to 30mm in length. In

addition, the highest stresses were observed in the crestal bone regions, regardless of the

implant length. This biotechnical analysis supports the opinion that longer implants are not

necessarily better. Instead, there is minimum implant length for each bone density,

depending on the width and design. The softer the bone, the greater the length suggested.

Crest Module Considerations:

The crest module of an implant body is the trans-osteal region from the implant

body and characterized as a region of highly concentrated mechanical stress. This region

of the implant is not ideally designed for load bearing, as evidenced by bone loss as a

common occurrence regardless of design or technique. Studies shown that mean marginal

bone loss of adjacent teeth recorded over the average time of examination (16 months) was

0.97± 1.46 mm was observed at upper lateral incisors facing a fixture in the canine or

central incisor regions". In fact, bone loss has been observed so often, many implant crest

modules are designed to reduce plaque accumulation once bone loss has occurred. A

smooth, parallel-sided crest m module will result in shear stresses in this region, making

maintenance of bone very difficult. An angled crest module of more than 20 degrees, with

a surface texture that increases bone contact, will impose a slight beneficial compressive

component to the contiguous bone and decrease the risk of bone loss

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 110


Fig : 9.5

The crest module of an implant should be slightly larger than the outer thread

diameter. Thus the crest module seats fully over the implant body osteotomy, providing a

deterrent for the ingress of bacteria or fibrous tissue. The seal created by the larger crest

module also provides for greater initial stability of the implant following placement,

especially in softer unprepared bone, as it compresses the region. The larger diameter also

increases surface area, which contributes to decreases in stress at the crestal region

compared with crest modules of smaller diameter.

Fig : 9.6

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 111


A polished collar of minimum height should be designed on the superior portion of

the crest module just below the prosthetic platform. A biologic width of 0.5mm has been

reported apical to the abutment-to-implant connection. A 0.5mm collar length provides for

a desirable smooth surface close to the peri-gingival area, while preserving the

biomechanical performance of the remaining portion of the crest module. Bone is subjected

to unnecessary and excessive shear loading in implants characterized by a longer polished

collar. Significant loss of crestal bone has been reported for implants with larger machined

(smooth) corona regions. This bone loss is attributed to the lack of effective mechanical

loading between the machined coronal region of the implant and the surrounding bone. This

clinical problem is reduced by a biomechanical design that minimizes the shear collar

surface area. It has been a common clinical observation that bone is often lost to the first

thread, regardless of the manufacturer type or design, after loading. Bone grows above the

threads during healing, but after prosthesis loading the bone loss is often observed .Yet, the

first thread is 1.2 mm below the platform of the Nobel Biocare implant, 2 mm below the

platform on the Steri-Oss design, and 3 mm on many Screw vent implant designs (Paragon).

The bone loss often stops at the first thread because, the first thread changes the shear force

of the crest module to a component of compressive force in which bone is strongest. The

studies indicated that some crestal bone loss occurred for both the threaded and partially

porous coated implants while no significant bone loss was seen with fully porous coated

implants'16. Instead of designing the crest module for shear, an improved design and/or

surface condition can reduce the crestal bone loss

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 112


Apical Design Considerations :

Most root form, implants are circular in cross-section. This permits a round drill to prepare

a round hole, precisely fitting the implant body. Round cross-sections, however, do not

resist torsional/shear forces when abutment screws are tightened or when free-standing,

single tooth implants receive a rotational (torsional) force. As a result, an antirotational

feature is incorporated, usually in the apical region of the implant body, with a hole or vent

being the most common design. In theory, bone can grow through the apical hole, and resist

torsional loads applied to the implant. The apical hole region may also increase the surface

area available to transmit compressive loads on the bone

A disadvantage of the apical hole occurs when the implant is placed through the

sinus floor or becomes exposed through a cortical plate. The apical hole may fill with mucus

and becomes a source of retrograde contamination or will likely fill with fibrous tissue.

Another antirotational feature of an implant body may be flat sides or grooves along the

body or apical region of the implant body. When bone grows against the flat or groove

regions, the bone are placed in compression with rotational loads. The apical end of each

implant should be flat rather than pointed. This allows for the entire length of the implant

to incorporate design features that maximize desired strain profiles. Additionally, if an

opposing cortical plate is perforated, a sharp, V-shaped apex may irritate or inflame the

soft tissues if any movement occurs (e.g., the inferior border of. the mandible).

Biomechanics of frameworks and misfit:

Frameworks:

The metal framework used in typical full-arch prosthesis with the Branemark system

can sometimes fracture in vivo. Unfortunately, no in-depth analyses of such fractures,

including case histories and explanations, exist. Nevertheless, it is possible to suggest some

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 113


reasons for these fractures, based on the biomechanical analyses presented so far. Fractures

have been observed to occur more towards die cantilever sections of the framework, for

example, just distal to the most distal implant. The fractures could be caused by two

mechanisms. One is outright overload of the cantilever by a single vertical bite force; the

distal portion of the prosthesis may bend like a cantilever beam and eventually fracture at

the root of the cantilever, where the stress is greatest. However, this mode of fracture is

unlikely in a reasonably-sized prosthesis made of a typical prosthetic alloy. The force

needed to induce fracture level stresses in the beam would be much larger than the typical

biting forces of a few hundred Newton. A more likely reason for prosthesis fracture is

metallurgical fatigue under cyclic biting loads. The stresses in the prosthesis caused by th e

cyclic forces of chewing day after day could produce stresses at the root of the cantilever,

which exceed the fatigue limit of the prosthetic alloy. To forestall such failures, the cross-

sectional areas of the framework near the root of the cantilever should be relatively

substantial, i.e. in the order of 3-6 mm on a size. This will help to reduce bending stresses

in this region because the stress varies with the square of the thickness of the beam and

linearly with its width.

Fig : 9.7

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 114


Gold screws and abutment screws mechanics:

When two parts are tightened together by a screw, this unit is called a screw joint.

The screw loosens only if outside forces trying to separate the parts are greater than the

force keeping them together. Forces attempting to disengage the parts are called joint

separating forces. The force keeping the parts together can be called the clamping force.

Joint-separating forces do not have to be eliminated to prevent screw loosening. The

separating forces must only remain below the threshold of the established clamping force.

If the joint does not open when a force is applied, the screw does not loosen. Therefore,

there are two primary factors involved in keeping implant screws tight;

(1) maximize clamping force,

(2) minimize joint-separating forces.

To achieve secure assemblies, screws should be tensioned to produce a clamping

force greater than the external force tending to separate the joint. In the design of a rigid

screw joint, the most important consideration from a functional standpoint is the initial

clamping force developed by tightening the screw. Joint strength is affected more by clamp

force than by tensile strength of the screws. Clamp load is usually proportional to tightening

torque.

Torque is a convenient, measurable means of developing desired tension. Too small

a torque may allow separation of the joint and result in screw fatigue failure or loosening.

Too large a torque may cause failure of the screw or stripping of the screw threads. Applied

torque develops a force within the screw called preload. Preload is the initial load in tension

on the screw. This tensile force on the screw develops a compressive clamping force

between the parts.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 115


Preload is determined by the following factors:

1. Applied torque.

2. Screw alloy.

3. Screw head design.

4. Abutment alloy.

5. Abutment surface.

In general, the more torque applied, the more preload generated. Two factors limit

the amount of torque that may be applied. The mechanical limit is the strength of the screw.

The amount of torque is also limited by how it is applied. Screwdrivers with larger handles

can generally apply more torque than those with small handles. A wrench can be used if

larger torques is needed.

In theory, the maximum preload is developed just before torsional fracture of the

screw occurs. Therefore, to increase preload and minimize the risk of screw fracture durin g

use, a safety margin is established. In simple terms, optimum tightening torques can be

calculated using 75% of the ultimate torque to failure values. In other words, the optimal

torque value can be calculated by tightening a screw until it fails; 75% of this value is the

optimum torque to place on the screw. In this manner, a significant clamping force can be

developed with minimum risk of screw fracture.

In industry, bigger screws are made to allow more torque than to be applied. In this

way, clamping force can be developed to resist nearly any joint separating force. It is not

that easy in the oral cavity. The size of the screws is limited by tooth size. The strength of

the bone implant interface is the biologic limit of applied torque.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 116


APPLICATION TO DENTAL IMPLANTS

The clinical reality is that implant restorations are continually subjected to joint -

separating forces. These forces include:

• Excursive contacts.

• Off-axis centric contacts.

• Angled abutments.

• Interproximal contacts.

• Wide occlusal table.

• Cantilever contacts.

• Non-passive framework.

Minimize Clinical Joint-Separating Forces:

The joint-separating forces can be greatly influenced by the moment arm through

which the force is applied. Excessive implant angles or prosthesis cantilevers can rapidly

magnify the centric contacts not aligned with the long axis of the implant and may increase

the joint-separating moment arm. Precision implant placement and treatment planning are

the first crucial step in maintaining tight implant screws.

Occlusion plays a primary role in keeping implant screws tight. Contacts in lateral

excursions act as separating forces and should be avoided whenever, possible. Remember,

however, that light lateral forces below the threshold of the clamping force do not cause

screw loosening. Therefore, minimal lateral guiding forces might be placed on anterior

implant restorations with, out adverse consequences.

The most commonly overlooked separating forces are off-axis centric contacts.

Normal centric contacts on molar cusp tips may exceed the clamping force threshold,

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 117


especially if the general occlusal force generated by the patient is large. This theory may

explain the high incidence of screw loosening in single-implants molars. Molar implant

screws should stay tight if the centric contacts can be directed in the long axis of the screw

and excursive contacts eliminated. Heavy interproximal contacts may also exert excessive

lateral force on an implant crown, resulting in screw loosening.

Attaching implants to natural teeth with a fixed partial denture can commonly lead

to loosing of screws in the implant abutment. The problem occurs because of mobility

differences between the two types of abutments. The implant is immobile relative to the

natural tooth, which can move within the limits of its periodontal ligament. Occlusal forces

on the natural tooth can have a cantilever effect on the implant, generating a maximum

resultant load up to two times the applied force. Much of this cantilever force is

concentrated at the joint between the implant crown and its abutment screw. It should not

be surprising that screws loosen in this clinical situation.

Likewise, screw-loosening incidents increase if a non passive framework is forced

to fit by tightening screws. The original framework applies joint-separating forces to the

system because it attempts to return to its original position. All non passive frameworks

should be sectioned and soldered to ensure passive fit.

Maximize Clinical Resistance to Joint Separation:

One possible advantage of the anti-rotational features used in dental is the resistance

they provide to joint-separating forces. The possibility that vertical walls engage between

the hexagon and the crown to resist applied force may explain the partial solution that these

devices provide. This occurrence would also explain why shorter hexes can allow some

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 118


screws to loosen under heavy loads. Studies shown the Modification of the single implant

system to use a flat headed screw reduced the loosening problem.

One of the simplest methods to ensure screw loosening is to make sure screws are

tight. The novice implant clinician often under-tightens the implant component. One study

suggests that the average torque placed with a screwdriver is only 11 Newton-cm (N-cm).

Most titanium components on the market can easily be tightened to twice that amount

without consequences. For torque levels greater than 20 N-cm, a torque wrench is usually

required. In reality, the optimum torque values for many of the larger-diameter implant

screws exceed the generally accepted limits of the bone-implant interface. Although

definitive torque removal values for the different implants have not been established in

humans, animal studies suggest that no greater than 30 to 35 N-cm of torque should be

applied to the bone-implant interface. In fact, the safest method of applying higher torque

values intraorally is to use a counter torque mechanism. If counter torque is applied to the

abutment as the screw is tightened, the net force at the bone interface should be zero.

Currently, torque levels in the 20- to 30-N-cm range are thought to provide significant

preload without risk to the bone interface. Studies suggested that 63 N-cm of torque could

be applied to the gold alloys screw before reaching the yield strength for the implant.

Titanium screws might also tolerate higher torque, to 39N-cm and still function with in the

materials elastic range.

Studies shown that there is a direct correlation between implant/ abutment

hexagonal rotational misfit and screw loosening. The better the matrix -to-patrix fit, the

more stable the screw joint. Less than 2 degrees of rotational freedom between the implant

external hexagonal extension and the abutment internal hexagonal recess resulted in the

most stable screw joint and the greatest resistance to screw loosening, with a mean of 6.7

million cycles and a 26% increase over the next larger abutment size. (1070 inch). Positive

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 119


hexagonal (External & Internal) engagement and elimination of rotational freedom resulted

in a stiffer screw joint that was substantially more resistant to screw loosening

Efforts have been made to reduce the rotational misfit between the coupling

hexagons to less than 4 degrees in the effort to reduce screw joint failure. One design

concept that uses an external hexagonal implant involves the 1.5 degrees tapered lock

developed. This effectively eliminated all rotational misfits. Another design concept that

uses an internal hexagonal recess with 45 degrees beveled with in the implant body and 1

degree tapered hexagonal extension on the abutment

The major clinical procedures necessary for tight implant screws are

summarized as follows:

1. Implants placed parallel to the forces of occlusion.

2. Restorations designed to minimize cantilever lengths.

3. Occlusion adjusted to direct forces in the long axis of the implant.

4. Antirotational feature engaged for single teeth.

5. Components tightened with 20-30 N-cm of torque, (unless specified by

manufacturer).

If screw loosening occurs, all potential contributing causes should be evaluated. The

clinician should pay particular attention to occlusal forces oblique to the implant long axis.

Interproximal contacts and framework fit should also be evaluated. Implant screws should

not be maximally tightened until joint-separating forces arc controlled.

We cannot focus only on eliminating loose screws; we must also eliminate the cause of

screw loosening.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 120


The danger for the patient lies in the fact that if the screws do not loosen, excessive

forces may be directed to more deleterious locations in the system. Proper implant

placement framework fit, and occlusal. Adjustment becomes even more important as screw

joints improve. If these fundamentals are not addressed, more stable screw connections

could result in fractured implant bodies or crestal bone loss. Loose screws should be seen

as a clinical symptom that may indicate that the forces are not appropriately balanced on a

particular implant restoration.

Framework misfit:

Most frameworks for full-arch prostheses are made using impressions, plaster

models, and casting techniques, etc. Despite every effort at precision, dimensional

inaccuracies inevitably occur in the final cast metal framework. Assuming that the misfit

is not too server, the framework may appear (at least by visual inspection) to fit well

‘passively, onto the abutments. However, there is increasing concern about the assessment

of ‘passive fit’ and its clinical significance.

A working definition of passive fit is suggested by a free body diagram. This shows

a framework for five abutments. Suppose four of the abutments match perfectly with the

gold cylinders in the framework. Assume that when each of the gold screws is torques down

onto the well-fitting abutments, the ideal preload of 300 N develops in each join. However,

suppose one of the five abutments does not fit well; note the gap (exaggerated, to make the

point) between one of the abutments and the framework in the figure Now, as the gold

screw is torque down to 10 N cm at the site of the gap, the tension which develop in the

gold screw and abutment screw will act on the framework, tending to bend it down toward

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 121


the abutment, diminishing the gap if the gap is small, it might be possible to close it

completely by such deformation of the framework.

However, if the gap is large, it may not be closed even when the nominal torque of

10-N cm is reached, in either case, the net effect is to apply a force on the framework at the

location of the misfit. This force can be considered as an ‘external’ force acting on the gram

work as per the Skalak Model; this means that the other four implants will be loaded by

virtue of the force on the framework at the side of misfit. Although this is a reasonable

theoretical explanation, clinical data on this effect are lacking.

SCIENTIFIC RATIONALE FOR IMPLANT AND PROSTHESIS SELECTION 122


TREATMENT PLANNING BASED ON
BIOMECHANICAL RISK FACTORS

The designing of an implant is very crucial and one of the most important aspect of an
implant therapy. There are four risk factors that need to be kept in mind while designing and
planning before an implant placement, which are as follows :
❖ Geometric risk factor.
❖ Occlusal risk factor.
❖ Bone implant risk factor.
❖ Technological risk factor.

I. GEOMETRIC RISK FACTORS: (can be related to the design and the anatomy)
1) When number of implants are less than number of root support :
➢ One implant replacing a molar – at high risk.
▪ 1 wide – platform implant and 2 regular implants.
➢ Two implants replacing 3 roots or more – at high risk
▪ 2 wide – platform implants
2) Wide – platform implants
➢ High risk – if used in very dense bone.
3) Implant connected to natural teeth .
4) Implants placed in a tripod configuration .
➢ Desired → counteract lateral loads
5) Presence of prosthetic extension.
6) Implants placed offset to the center of the prosthesis → in tripod arrangement, offset is
favorable
7) Excessive height of the restoration.

Fig: 10.1

TREATMENT PLANNING BASED ON BIOMECHANICAL RISK FACTORS 123


II. OCCLUSAL RISK FACTORS:
❖ Force intensity and parafunctional habit.
❖ Presence of lateral occlusal contact.
❖ Centric contact in light occlusion.
❖ Lateral contact in heavy occlusion.
❖ Contact at central fossa.
❖ Low inclination of cusp.
❖ Reduced size of occlusal table.

Fig : 10.2

III. BONE IMPLANT RISK FACTORS:


o Dependence on newly formed bone.
o Absence of good initial stability.
o Proper healing time before loading.

IV. TECHNOLOGICAL RISK FACTORS:


▪ Lack of prosthetic fit and cemented prostheses.
▪ Not following proven and standardized protocols.
▪ Not following instrumentation with stable and pre-defined values .

TREATMENT PLANNING BASED ON BIOMECHANICAL RISK FACTORS 124


CONCLUSION

Implant therapy has gained popularity over the years has become quite common. A
complete and thorough understanding of its structure and its behaviours and response under
function is very much important. It can be stated that the most important step in an implant
treatment is the planning stage, so it is paramount that a complete scientific knowledge of the
implant with its planned superstructure and its expected response to its surrounding tissue after
placement is known to the clinician., so that a better longevity can be expected.

Recently there are many documented cases of failed implants just because of the poor
planning which that maybe have been underwent before the implant placement and the scarse
knowledge about the biomechanics of the implant after being placed. It can be said that
although osseointegrated dental implants have been used for over three decades, an ideal
implant has not been designed so far according to the biomechanical needs of the bone. In
essence, it is still unknown to us whether it is the implant design or loading history or both in
respect to the implant that rule bone adaptation mechanism.

Biomechanical considerations in implant dentistry to a large extent follow simple


mechanical rates, based on leverage principles and the principles of initial implant stabilization.
The control of forces may be more important than the very nature of implants for the
maintenance of marginal bone. Thus, the sound knowledge of biomechanics will possibly
minimize the situations in which there maybe chances of overloading and generation of
unwanted stresses and loads action on the implant or its superstructure which is very crucial
and absolute necessary for long-term success of the dental implant.

CONCLUSION 125
BIBLIOGRAPHY

1. Albrektsson T, Branemark P A. Osseointegrated titanium implants: requirements for


ensuring a long-lasting, direct bone-to-implant anchorage in man. Acta Orthopaedica
1981; 52(2): 155–170.
2. Cook SD, Klawitter JJ, Weinstein AM. A model for the implant-bone interface
characteristics of porous dental implants. J Dent Res 1982; 61(8): 1006–1009.
3. Skalak R. Biomechanical considerations in osseointegrated prostheses. J Prosthet Dent
1983; 49: 843-50.
4. Borchers L, Reichart P. Three dimensional stress distribution around a dental implant
at different stages of interface development. J Dent Res 1983; 62(2): 155-159.
5. Carter D R. The relationship between in viuo strains and cortical bone remodelling.
Crit Rev Biomed Engr 1983; 8: 1-10.
6. Albrektsson T, Zarb G, Worthington P, Eriksson A R. The long term efficiency of
currently used dental implants: a review and proposed criteria for success. Int J Oral
Maxillofac Implants 1986; 1: 11-25.
7. Hayes W C. Bone mechanics: from tissue mechanical properties to an assessment of
structural behaviour. In Frontiers of Biomechanics 1986; 196-209.
8. Frost HM. Bone mass and the Mechanostat: A proposal Anat Rec 1987; 219: 1-9.
9. Kinni M E. Hokama S N. Force Transfer by Osseointegration Implant Devices.
International Journal of Oral & Maxillofacial Implants 1987; 11-17.
10. Rieger M R, Adams W K, Kinzel G L, Brose M O. Finite element analysis of bone-
adapted and bone-bonded endosseous implants. J Prosthet Dent 1989; 62(4): 436–440.
11. Maeda Y, Wood W W. Finite element method simulation of bone resorption beneath a
complete denture. J Dent Res 1989; 68: 1370-1373.
12. Setz J, Kramer A. Complete dentures fixed on dental implants: chewing patterns and
implant stress. Int J Oral Maxillofac Implants 1989; 4(2): 107-111.
13. Zarb G A, Schmitt A. The longitudinal clinical effectiveness of osseointegrated
implants: the Toronto stud,. part III: problems and complications encountered. J
Prosthet Dent 1990; 64(2): 185–194.
14. Williams K R, Watson C J, Murphy W M. Finite element analysis of fixed prostheses
attached to osseointegrated implants. Quintessence Int 1990; 21: 563–570.

BIBLIOGRAPHY viii
15. Cowin S C. Bone mechanics, 1990 (CRC Press, Boca Raton,Florida).
16. Van Rossen I P, Braak L H, Putter C, Groot K. Stress absorbing elements in dental
implants. J Prosthet Dent 1990; 64(2): 198–205.
17. Takuma M, Tsutsumi S, Kurokawa F, Takashima F, Miyauchi S. The influence of
materials difference on stress distribution and bone remodelling around alumina and
titanium dental implants. Journal of Osaka University Dental School 1990; 30: 96-102.
18. Brunski J B. Biomechanical factors affecting the bone dental implant interface, Clinical
Materials 1992; 10: 153–201.
19. Quirynen M, Naert I. Fixture design and overload influence marginal bone loss and
fixture success in the Branemark system. Clin Oral Implants Res 1992; 3: 104-111.
20. Clift S E, Fisher J. Finite element stress and strain analysis of the bone surrounding a
dental implant: effect of variations in bone modulus. Journal of Engineering in
Medicine 1992; 233-241.
21. Meijer H J, Starmans F J, Bosman F, Steen W H. A comparison of three finite element
models of an edentulous mandible provided with implants. J Oral Rehabil 1993; 20:
147-157.
22. Frost H M. Wolff’s Law and bone’s structural adaptations to mechanical usage: an
overview for clinicians. Angle Orthod 1994; 64: 175–188.
23. Rodriguez A, Aquilino S A. Cantilever and Implant Biomechanics: A Review of the
Literature. Journal of Prosthodontics 1994; 3(2): 114-119.
24. Vaillan Court H, Pilliar R M, mccammond d. Finite element analysis of crestal bone
loss around porous-coated dental implants. J Appl Biomater 1995; 6: 267–282.
25. Rangert B, PallacI P. Optimal implant positioning and soft tissue management for the
Branemark System.Chicago: Quintessence Publishing Co Inc 1995; 21-39.
26. Mericske-Stern R, Assal P, Mericske E, Burgin W. Occlusal force and oral sensibility
measured in partially edentulous patients with ITI implants. Int J Oral Maxillofac
Implants 1995; 10: 345–354.
27. Melo C, Matsushita Y, Koyano K, Hirowatari H. Comparative stress analyses of fixed
free-end osseointegrated prostheses using the finite element method. J Oral Implantol
1995; 21: 290-294.
28. Haack J E, Sakaguich R L, Sun T, Coffey J P. Elongation and preload stress in dental
implant abutment screws. Int J Oral Maxillofac Implants 1995; 10: 529-536.

BIBLIOGRAPHY ix
29. Benzing U R, Gall H, Weber H. Biomechanical aspects of two different implant-
prosthetic concepts for edentulous maxillae. Int J Oral Maxillofac Implants 1995; 10:
188-198.
30. Patterson E A, Burguete R L, Hue Thoi M, Johns R B. Distribution of load in an oral
prosthesis system: An in vitro study. Int J Oral Maxillofac Implants 1995; 10: 552–560.
31. Akpinar I, Demirel F, Parnas L, Sahin S. A comparison of stress and strain distribution
characteristics of two different rigid implant designs for distal-extension fixed
prostheses. Quintessence Int 1996; 27(1): 11–17.
32. 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.
33. Teixeira E R, Sato Y. A comparative evaluation of mandibular finite element models
with different lengths and elements for implant biomechanics. Journal of Oral
Rehabilitation 1998; 25: 299–303.
34. Lai H, Zhang F, Zhang B. Influence of percentage of osseointegration on stress
distribution around dental implants. The Chinese Journal Of Dental Research 1998;
1(3) : 7-11.
35. Zhou X, Zhao Z, Zhao M, Fan Y. The boundary design of mandibular model by means
of the three-dimensional finite element method. West China J Stomatol 1999; 17: 1-6.
36. Brunski J, Puleo D, Nanci A. Biomaterials and biomechanics of oral and maxillofacial
implants: current status and future developments. Int J Oral Maxillofac Implants 2000
; 15(1) : 15–46.
37. Çiftçi Y, Canay S. The effect of veneering materials on stress distribution in implant-
supported fixed prosthetic restorations. Int J Oral Maxillofac Implants 2000; 15(4):
571–582.
38. O’ Mahony A, Bowes Q. Stress distribution in the single-unit osseointegrated dental
implant: finite element analyses of axial and off-axial loading. Implant Dent 2000; 9(3):
207-218.
39. Akca K. Finite Element Stress Analysis of the Influence of Staggered Versus Straight
Placement of Dental Implants. The International Journal Of Oral And Maxillofacial
Implants 2001; 16(5): 722-723.
40. Geng J P, Tan B C. Application of finite element analysis in implant dentistry: A review
of the literature. J Prosthet Dent 2001; 85(6): 585-598.

BIBLIOGRAPHY x
41. Pierrisnard L, Hure G, Barquins M, Chappard D. Two dental implants designed for
immediate loading: a finite element analysis. Int J Oral Maxillofac Implants 2002;
17(3): 353-362.
42. Natali A. Dental Biomechanics. New York: CRC Press. 2003.
43. Ishigaki S, Nakano T, Yamada S, Nakamura T. Biomechanical stress in bone
surrounding an implant under simulated chewing. Clin Oral Impl Res 2003; 14: 97-102.
44. Cruz M, Wassall T. Three-dimensional Finite Element Stress Analysis of a Cuneiform-
Geometry Implant. The International Journal of Oral & Maxillofacial Implants 2003;
18(5): 675-686.
45. Himmlova L, Dostalova T, Kacovsky A. Influence of implant length and diameter on
stress distribution: a finite element analysis. J Prosthet Dent 2004; 91(1): 20-25.
46. Lemons J E. Biomaterials, biomechanics, tissue healing, and immediate function dental
implants. J Oral Implantol 2004; 5: 318-324.
47. Cehreli M C, Duyck J, Cooman M, Puers R, Naert I. Implant design and interface force
transfer: A photoelastic and strain-gauge analysis. Clin Oral Implant Res 2004; 15: 249-
257.
48. Kitamura E, Stegaroiu R, Nomura S, Miyakawa O. Biomechanical aspects of marginal
bone resorption around osseointegrated implants: Considerations based on a three-
dimensional finite element analysis. Clin Oral Implant Res 2004; 15: 401-412.
49. Geng J P , Beng W X, Tan K B C, Liu G R. Finite element analysis of an osseointegrated
stepped screw dental implant. J Oral Implantol 2004; 4: 223-233.
50. Alkan I, Sertgöz A, Ekici B. Influence of occlusal forces on stress distribution in
preloaded dental implant screws. J Prosthet Dent 2004; 91: 319-325.
51. Eskitascioglu G, Usumez A, Sevimay M, Soykan E, Unsal E. The influence of occlusal
loading location on stresses transferred to implant-supported prostheses and supporting
bone: A three-dimensional finite element study. J Prosthet Dent 2004; 91(2): 144-150.
52. Satoh T, Maeda Y, Komiyama Y. Biomechanical rationale for intentionally inclined
implants in the posterior mandible using 3D finite element analysis. Int J Oral
Maxillofac Implants 2005; 20: 533-539.
53. Jingade R R K, Rudraprasad I V. Biomechanics of dental implants: A FEM study. The
Journal of Indian Prosthodontic Society 2005; 5(1): 18-22.
54. Kitagawa T, Tanimoto Y, Nemoto K, Aida M. Influence of cortical bone quality on
stress distribution in bone around dental implant. Dent Mater J 2005; 24: 219-224.

BIBLIOGRAPHY xi
55. Kim Y, Misch C.E, Wang H L. Occlusal considerations in implant therapy: clinical
guidelines with biomechanical rationale. Clinical Oral Implants Research 2005; 16: 26–
35.
56. Noguerol B, Munoz R. Early implant failure: prognostic capacity of periotests:
retrospective study of a large sample. Clin Oral Impl Res 2006; 17: 459–464.
57. Misch C E. Contemporary Implant Dentistry. Maryland Heights, MO : Elsevier Health
Sciences 2008.
58. Skerry T M. The response of bone to mechanical loading and disuse: fundamental
principles and influences on osteoblast/osteocyte homeostasis. Arch Biochem Biophys.
2008; 473: 117–123.
59. Baggi L, Cappelloni I, Maceri F, Vairo G. Stress-based performance evaluation of
osseointegrated dental implants by finite-element simulation. Simulation Modelling
Practice and Theory 2008; 16 (8) 971–987.
60. Chen Y, Kuan C L. Implant occlusion: biomechanical considerations for implant-
supported prostheses. J Dent Sci 2008; 3(2): 65-75.
61. Kong L, Hu K, Li D, Song Y, Yang J, Wu Z, Liu B. Evaluation of the cylinder implant
thread height and width: a 3-dimensional finite element analysis. Int J Oral Maxillofac
Implants 2008; 23: 65-74.
62. Bergkvist G, Simonsson K. A finite element analysis of stress distribution in bone tissue
surrounding uncoupled or splinted dental implants. Clin Implant Dent Relat Res 2008;
10(1): 40-46.
63. Bellini C M, Romeo D. Comparison of Tilted Versus Nontilted Implant-Supported
Prosthetic Designs for the Restoration of the Edentuous Mandible: A Biomechanical
Study. Int J Oral Maxillofac Implants 2009; 24: 511–517.
64. Rubo J H. Biomechanical studies in implant prosthodontics. Implant News 2010; 7:
139-44.
65. Faegh S, Müftü S. Load transfer along the bone-dental implant interface. Journal of
Biomechanics 2010; 43: 1761-1770.
66. N Djebbar, Serier B. Analysis of the effect of load direction on the stress distribution
in dental implant. Materials And Design 2010; 31(4): 2097-2101.
67. Hasan I, Bourauel C, Keilig L. The influence of implant number and abutment design
on the biomechanical behaviour of bone for an implant-supported fixed prosthesis: a

BIBLIOGRAPHY xii
finite element study in the upper anterior region. Computer Methods in Biomechanics
and Biomedical Engineering 2011; 14(12) 1113–1116.
68. Kim K S, Kim Y L. Biomechanical comparison of axial and tilted implants for
mandibular full-arch fixed prostheses. The International Journal of Oral &
Maxillofacial Implants 2011; 26(5): 976-986.
69. Lee W T, Koak J Y, Lim Y J, Kim S K, Kwon H B, Kim M J. Stress shielding and
fatigue limits of poly-ether-ether-ketone dental implants. J Biomed Mater Res & Appl
Biomater. 2012; 100: 1044–1052.
70. Ozkir S E, Terzioglu H. Macro design effects on stress distribution around implants: A
photoelastic stress analysis. Indian Journal of Dental Research. 2012; 23(5): 603-608.
71. Lan T H, Du J K, Pan C Y, Lee H E, Chung W H. Biomechanical analysis of alveolar
bone stress around implants with different thread designs and pitches in the mandibular
molar area. Clin Oral Investig 2012; 16: 363-369.
72. Woo Taek Lee, Journal of Biomedical Materials Research Part B Applied
Biomaterials 2012; 100(4):1044-1052.
73. Ozkir S E. Macro design effects on stress distribution around implants: A photoelastic
stress analysis. Indian Journal of Dental Research 2012; 23(5): 603-607.
74. Amid R, Raoofi S, Kadkhodazadeh M, Movahhedi M R, Khademi M. Effect of micro-
thread design of dental implants on stress and strain patterns: a three-dimensional finite
element analysis. Biomed Tech (Berl) 2013; 58: 457-467.
75. Chang H S, Chen Y C. Stress distribution of two commercial dental implant systems:
A three-dimensional finite element analysis. Journal of Dental Sciences 2013: 1-11.
76. Mathieu V, Vayron R, Richard G. Biomechanical determinants of the stability of dental
implants: influence of the bone-implant interface properties. J Biomech. 2014; 47: 3–
13.
77. Gehrke S A , Santos Vianna M S. Influence of bone insertion level of the implant on
the fracture strength of different connection designs: an in vitro study. Clinical Oral
Investigations 2014; 18.(3): 715–720.
78. Negri B, Calvoguirado J L, Maté Sánchez Deval J E, Delgado Ruíz R A, Peri-implant
tissue reactions to immediate nonocclusal loaded implants with different collar design:
an experimental study in dogs. Clin Oral Implants Res 2014; 25: 54-63.

BIBLIOGRAPHY xiii
79. Karaçali O. Material fatigue research for zirconia ceramic dental implant: a
comparative laboratory and simulation study in dentistry. ACTA PHYSICA
POLONICA 2015; 127 (4); 1195-1198.
80. Dhatrak P, Shirsat U. Fatigue life prediction of commercial dental implants based on
biomechanical parameters: a review. Journal of Materials Science & Surface
Engineering 2015; 3 (2): 221-226.
81. Mannarino F S, Carvalho R S. Analysis of the distribution of stress and deformation
in single implant-supported prosthetic units in implants of different diameters. Rev
Odontol UNESP 2016; 45(5): 247-252.
82. Privado M P, Prados Frutos J C. Long-Term Fatigue and Its Probability of Failure
Applied to Dental Implants. BioMed Research International 2016; 1-8.
83. Yadav P, Tahir M, Shetty P, Saini V, Prajapati D. Implant design and stress distribution.
Int J Oral Implantol Clin Res 2016; 7(2): 34-39.
84. Yazicioglu D, Bayram B. Stress Distribution on Short Implants at Maxillary Posterior
Alveolar Bone Model With Different Bone-to-Implant Contact Ratio: Finite Element
Analysis. Journal of Oral Implantology 2016; 42(1): 27-34.
85. Korabi R, Shemtov-Yona K, Rittel D. On stress/strain shielding and the material
stiffness paradigm for dental implants. Clin Implant Dent Relat Res 2017; 1–9.
86. Hongyi F, Xueqi G, Zhuoli Z. Evaluation of dental implant fatigue performance under
loading conditions in two kinds of physiological environment. Int J Clin Exp Med.2017;
10(4): 6369-6377.
87. Manikyamba Y J B, Sajjan S M C. Implant thread designs : An overview. Trends In
Prosthodontics and Dental Implantology 2017; 8: 11- 20.
88. Geramizadeh M, Katoozian H. Comparison of finite element results with photoelastic
stress analysis around dental implants with different threads. Dent Med Probl 2018;
55(1): 17–22.
89. Jaros O A, De Carvalho G A, Franco A B, Kreve S. Biomechanical behavior of an
implant system using polyether ether ketone bar: Finite element analysis. J Int Soc
Prevent Communit Dent 2018; 8: 446-450.

BIBLIOGRAPHY xiv

You might also like