Sunteți pe pagina 1din 21

American Journal of ORTHODONTICS

Volume 80, Number 5 November, 1981

ORIGINAL ARTICLES

Understanding, planning, and


managing tooth movement:
Orthodontic force system Dr. Hocevar

theory
Richard A. Hocevar, D.M.D.*
Dunedin, New Zealand

A simple force may cause translation andior rotation of the tooth upon which it acts.
The closer the line offorce to the tooth’s center of resistance, the greater is the
proportion of translation relative to rotation, and vice versa. When a tooth is subjected
to a tipping moment, strain is concentrated in the areas of the alveolar crest and root
apex. This may explain how a light force can tip a tooth readily, while translation,
involving a more even distribution of strain throughout the length of the root, requires
more force with little or no moment. It is d@cult to sustain a desired combination of
rotational and translational movement over any signtjicant distance. Couples induce
rotation only. Couples and simple forces can be used in combinations to effect
movements that are not possible with either one alone. Molars provide anchorage for
incisor intrusion. This “vertical anchorage” can be reinforced by elastics producing
moments opposite those produced by arch wire anchor bends. Calculation of the
elastics’ contribution provides an indication of the intrusive force that muy be applied
to incisors without reciprocal molar tipping. The common use of the term center of
rotation in orthodontics may be inappropriate andior confusing. It may be more
accurate and straightforward to view most tooth movements as combinations of linear
translation with rotation about the center of resistance.

Key words: Biomechanics, force systems, Begg technique, tooth movement,


anchorage

*Lecturer in Orthodontics, University of Otago Dental School.

0002-9416/81/l 10457+21$02.10/O 0 1981 The C. V. Mosby Co 457


There are many aspects of orthodontic biomechanics that are worthy and in
need of clarification and a number of subtleties of treatment with the Begg technique
which deserve more attention than they have been allotted in the many descriptions of this
technique. The present article will examine and emphasize certain mechanical principles
which, while they may appear to be merely minor details or abstract theoretical concepts,
could be quite important to the success of treatment. It may be regarded as a working
hypothesis for the treatment of orthodontic patients, a sort of “artist’s conception” of the
workings of force systems, intended to illustrate and explain in a simplistic fashion the
manner in which forces applied in orthodontic treatment bring about specitic predictable
tooth movements. Subsequent articles will extend the foregoing approach, portraying
applications of theoretical principles to orthodontic technique and presenting technique
refinements as logical, practical derivatives of these principles. I make no pretense of
presenting rigorous scientific investigation. A helpful explanation of the control of tooth
movement is offered, plus a number of suggestions that can result in more efficient and
effective treatment. While this article is addressed specifically to Begg practitioners, many
of the principles and suggestions herein are equally relevant to other fixed-appliance
techniques.

The relation between a line of force and resulting tooth movement


The spatial relationships between forces and roots determine how teeth will move.
This discussion will cover only situations in which forces are applied via “single-point”
nonrigid contacts with the teeth (as in “Begg brackets”),’ ’ except where specifically
stated otherwise.
Let us hypothesize an idealized system in which a tooth is thought of as a two-
dimensional body having its mass uniformly distributed throughout its root (the crown and
bracket being massless) and lying on a frictionless plane. The tooth’s mass may be
considered as being concentrated in a single point in the center of the root, the center of
mass (CM). Any force applied at the bracket would cause translation of the center of mass
along a line parallel to the line of force, the acceleration being proportional to force
magnitude. Forces directed along the bracket-CM line would yield translation only. Any
force not along that line would also create a moment resulting in rotation of the tooth about
CM (which would be moving). The magnitude of the moment produced, and thus the rate
of induction of rotation, is directly proportional to the perpendicular distance from CM to
the line of force (the “moment arm”) and to the force magnitude (M = F X d, where M
is the moment, F is the force, and d is the moment arm). Thus, a force passing very close
to CM would cause very little rotation relative to translation, whereas a force perpendicu-
lar to the bracket-CM line at the bracket would yield a much greater degree of rotation
relative to translation.
An actual tooth in a patient represents a somewhat different situation. We can still
enjoy the luxury of considering the tooth as being two dimensional and resistance to
movement as being uniform over the root surface, 6-H although these assumptions are, of
course, simplifications which may decrease the degree of accuracy of the conceptualiza-
tion. Instead of the inertia of a mass on a frictionless plane, bone and periodontal ligament
provide resistance to movement, and the concept of center of mass is replaced by that of
the center of resistance (CR). CR is assumed to be analogous to CM and to correspond,
approximately, with the geometric center of the root. It is conceivable, however, that the
Volume 80
Understanding, planning, and managing tooth movement 459
Number 5

response of the supporting tissues may not be uniform over the area of a root, consistent
among different types of tooth movement, or constant throughout movement. The tissues
themselves must undergo varying catabolic and anabolic changes, so that CR may not be
always in the exact center of the root.
The situation in which the line of force passes through CR is still straightforward: the
tooth translates along the line of force. However, when the force does not pass through
CR, the real response may differ from the hypothetical one. Movement of a tooth through
bone depends upon a biologic response whose mechanism is apparently sensitive to
differentials in pressure and tension within the periodontal ligament, electrical potential,
and/or deformation or “bending” of the bone,g and which is time dependent. These
concepts of “pressure, ” “bone bending,” etc. may be represented best by the word
strain, the term used in physics for deformation caused by a stress, or force.
A tooth can move only when and where physiologic processes have created space for it
by bone resorption. It is not known whether, in absolute terms, a threshold or minimal
force level is required to induce tooth movement. “3 I’ Forces as light as 2 grams have been
found to tip teeth.12 However, it is known that while tipping occurs readily and rapidly
with very light forces, translation occurs reluctantly and slowly, requiring much heavier
forces. l3 Forces of 15 to 20 Gm. are highly effective in tipping teeth,‘O and rates of crown
movement on the order of 1 mm. per week can be attained,14 whereas even forces of 100
to 1,000 Gm. tend to produce translation at less than one tenth this rate.‘”
When a moment created by a light force is applied to a tooth, the strain is concentrated
in the areas of the alveolar crest and apex and is adequate to induce a biologic response
yielding rapid movement of these portions of the root and of the crown. The strain around
the middle third of the root is minimal, so that translation of CR is insignificant on a
clinically relevant time scale; it is completely outstripped and overshadowed by the
tipping. If the moment were decreased by directing the force very close to CR, then there
would be significant translation relative to tipping, but the movement would be very slow
because the strain would be relatively evenly distributed throughout the socket area and
would be minimal at any particular point. An attempt to increase the absolute rate of
translation would require a much greater force to produce sufficient strain throughout the
socket area to stimulate enough cellular activity to allow the root to move bodily.
This is a plausible explanation for the clinical impression that teeth can be tipped, but
not translated, easily and quickly with light forces, and for the anchorage phenomenon of
“differential response to force”’ employed in the Begg technique. Light force, just
sufficient to produce rapid tipping if applied in such a way as to concentrate strain toward
apex and alveolar crest, may cause negligible movement in a clinically relevant time
period if delivered so as to spread the strain evenly over a large socket area so that the
strain is not sufficient at any point to incite enough osteoclastic activity to allow the tooth
to move. Thus can anterior teeth be tipped posteriorly readily by a force whose reaction is
too light to protract posterior teeth that are not allowed to tip.
Bodily retraction of the anterior teeth would require more (in magnitude and/or dura-
tion) force, whose reaction might protract the posterior teeth. If a heavy force were
employed, unless its moment arm on the anterior teeth were kept very small, it might
produce extreme strain at the alveolar crest and apex, perhaps even inhibiting the biologic
processes required for tooth movement,4’ I13 I43 l6 necessitating undermining resorption
and inducing pathosis.‘0s 16-lx Th’IS d’iscussion will be concerned only with light forces.
Fig. 1. Burstone proposed that M/F = 0.066 h*/y. Here F is a simple force; that is, it may be the re-
sultant of several forces produced by arch wire, elastics, and/or headgear, but is delivered to the tooth
at a single-point nonrigid contact involving no couples. M is the moment, = F x d; h is root length, say
Fxd
12 mm.; y is the distance of the “center of rotation” from CR. Thus, ~ = 0.066 x 144 mm.*ly mm.
F
which yields approximately d = 10/y or y = 10/d. Therefore, if d 2 5 mm., y s 2 mm., and if d 2 3
mm., y % 3.33 mm. That is, even when the line of force passes within 3 mm. of CR, the tooth movement
may be thought of as resembling a rotation around a point very near CR, and the further the line of force
from CR, the more closely the movement approximates pure rotation about CR.

Fig. 2. A line of force passing from the bracket through CR effects translation.

It is true that simple forces of a given magnitude in any direction would have equal
tendencies to translate the CR of a tooth in the direction of the force; however, for any
forces other than those traversing the root area, the actual translation would be minimal or
negligible; movement would be virtually limited to rotation about CR.‘, H, I1 Burstone’s
equation relating moment, force, root length, and location of “center of rotation, “I9 can
be used to show, roughly, the sort of effect that the length of the moment arm has upon the
type of tooth movement induced by a force (Fig. 1).
Volume 80
Number 5 Understanding, planning, and managing tooth movement 461

A C D

Fig. 3. Consideration of its effect on the periodontal tissues shows why a horizontal force produces
virtually pure tipping. Lines perpendicular or oblique to the root surface represent “tension” in the
penudontal tissues, and lines parallel to the surface represent “compression.” Longer and heavier
lines represent greater “tension” and “compression,” respectively. The neutral zone between “tension”
and “compression” is slightly further apical in D than it would be for pure tipping (B); the coronal portion
of the tooth moves slightly more than the apical portion; there is a scintilla of horizontal translation. (See
text.)

A B C D
Fig. 4. The effect of a vertical force. (See legend for Fig. 3 and text.)

A B C D
Fig. 5. See legend for Fig. 3 and text. As this line of force passes closer to CR, the moment is not so
great, and the rotational aspect of the movement is less predominant.
Fig. 6. (See text.)

Fig. 7. (See text.)

Fig. 8. With the tooth in its initial position (clear) the line of force passes close to CR; movement is a
combination of tipping and translation. If the force direction stays the same, the distance from the line of
force to CR increases as the tooth moves, a greater moment is produced, and rotation camp&es an
increasing proportion of the movement.
Volume 80
Number 5
Understanding, plunning, and managing tooth movement 463

F3

Fig. 9. (See text.)

The situation is illustrated schematically in Figs. 2 to 8. In Fig. 2 the line of force


intersects CR. There is no torque, and the tooth translates along the line of force if the
strain is of sufficient magnitude to initiate the necessary cellular response. Fig. 3 illus-
trates a horizontal force, acting at roughly 45 degrees to the bracket-CR line. Fig. 3, A
shows the effect of the translatory aspect of the force on the periodontal tissues (note the
even distribution of strain); Fig. 3, B shows the rotational aspect (strain concentrated at
alveolar crest and apex); Fig. 3, C shows the summation of the two effects; and Fig. 3, D
shows the net result (with strain concentration as in Fig. 3, B), virtually pure rotation, or
tipping. Fig. 4 similarly illustrates the case in which the force is directed vertically
upward. In Fig. 5 the moment is reduced; consequently, the movement of the tooth is a
combination of translation of CR in the direction of the force and rotation about CR.
Forces through the wide range of directions through the arcs in Fig. 6 will produce
virtually simple tipping. Forces through the cross-hatched areas might initially yield
minimal translation as well as rotation. Forces passing through the directions shown in
Fig. 7 will produce translation combined with rotation at a rate which is continuously
accelerating because, if the force is maintained constant, as the tooth moves the line of
force passes further from CR and the length of the moment arm increases (Fig. 8).
This concept is the bane of any attempt at any sustained movement other than simple
tipping by means of a single, freely pivotting contact point on the tooth. If one attempts to
direct the line of force, which obviously cannot be determined accurately, very close to
CR, there is risk of its actually passing on the wrong side of CR and thus inducing rotation
opposite that desired. If one attempts to ensure that the line of force will be on the correct
side of CR to produce rotation in the desired direction, the result is likely to be almost pure
tipping. The greater the distance of combined rotation-translational tooth movement
being attempted, the more difficult it is to plan and maintain a force system to achieve
it. H. 11
Couples and torque. A couple consists of two forces of equal magnitude acting in
parallel but opposite directions and having different points of application. Since the forces
composing a couple act in opposite directions, it induces no translation. It does, however,
produce a moment (MC) inducing rotation about CR, whose magnitude is the product of
the distance between the lines of force composing the couple (dc) and their magnitude
(MC = F X dc). (MC is the sum of the moments that each of the forces would produce on
:

F2

C D
Fig. 10. A, If the moments produced by the couple, Fz and F, and the force, F,, are equal, (that is,
F,d = F, d,), then the tooth will be translated in the direction of F,. (The illustration portrays only force
directions, not magnitudes.) B, lf F3d, > F,d, there will be translation (CR has moved lingually) plus
clockwise rotation. C, If F, d, >> F,d, there will be less translation and more rotation; ultimately
F,d, >>> F,d would produce pure rotation about CR. D, If F,d > F,dc, there will be lingual tipping and
some translation along F,. In each case there would also be a downward force (not shown; see Figs. 11
and 12, D) whose magnitude would ba proportional to the moment produced by the couple.

its own.) In Fig. 9 F, = F, = F, and d = dc. The moments produced by the simple force,
F, , M = F, x d, and the couple, F, and Fa, MC = F, x dc, are equal in magnitude but
opposite in direction, so the total moment on the tooth is 0. Therefore, the tooth translates
in the direction of F, but does not rotate.
This sort of balancing of couples and forces is necessary for any controlled movement
of roots, whether by auxiliaries or rectangular wires. Real lingual root torque for maxillary
incisors requires both a couple to produce the rotation and a force to prevent the brackets’
moving anteriorly, that is, to translate CR posteriorly. The force (F, in Fig. 10) may be
derived externally, as with headgear or Class II elastics, or generated from within the arch;
Understanding, planning, and managing tooth movement 466

Fig. 11. Begg torquing arch (left) and rectangular wire in edgewise bracket (right), passive. Activation
(dashed arrows) tends to extrude anterior teeth and, reciprocally, intrude posterior teeth. Solid arrows
indicate the systems of forces produced by the activated wires.

if the arch wire is “cinched back, “4 or if anterior and posterior teeth are ligated together, l the
posterior teeth provide resistance to movement in reaction to the tendency for the incisor
crowns to move forward under the influence of the couple.
Unfortunately, many mechanisms for torquing incisor roots lingually also induce
incisor extrusion and molar intrusion (Fig. 11). This means that it is not practical to
intrude and retract the roots of retroclined incisors simultaneously. Class II, Division 2
cases often require that bite opening be accomplished first, followed by root retraction.
Conversely, all mechanisms for producing lingual crown torque also induce incisor
intrusion and molar extrusion. Thus, there are two possible approaches to consider for
treatment of Class II, Division 1 cases with proclined incisors when upper incisor intru-
sion is desired: (1) One can retract the incisor crowns, either by simple tipping (Fig. 12, A)
or by combined tipping and translation (Fig. 12, B), and then, after the incisors are upright
or somewhat retroclined, intrude them by adjusting the force vertically to pass through or
near CR (Fig. 12, C); after intrusion beyond the desired level, lingual root torque is
applied as necessary (Fig. 12, D). (2) One can apply a force and couple as in Fig. 13, the
couple producing a moment of lingual crown torque greater than the moment of facial
crown torque produced by the force, and, after intrusion beyond the desired level, apply
lingual root torque as necessary. The hrst approach is similar to typical Begg treatment,
except the addition of a phase in which intrusion is sought after retroclination is
suggested. This approach demands horizontal anchorage throughout, used moderately in
the first phase, lightly in the second, and heavily in the third. In theory, the second
approach (Fig. 13) should produce in its first-phase upward and backward translation
(along the line of force) plus retroclination (induced by the couple) for minimal anchorage
taxation. In fact, the lingual crown torque should induce posterior crown tipping of the
entire maxillary dentition and augmentation of incisor intrusion for the “price” of some
upper molar extrusion. This approach is currently under clinical investigation (Fig. 14).
Note that with proclined incisors it is not possible to achieve simultaneously both
significant intrusion and retroclination with only a single constant line of force. Forces
such as those shown in Fig. 12, A and B would retrocline incisors, but their directions are
C D
Fig. 12. A proclined incisor is retracted in the Begg technique by either simple tipping (A) or tipping plus
a little translation (6). Note in B that even though CR moves slightly upward, the incisal edge moves
downward. Once the tooth had been retroclined, modifying the force system could produce some
intrusion (C). “Lingual root torque” (D) corrects the inclination; notice how much downward movement
of CR, the measure of true extrusion, occurs with only slight movement of bracket and incisal edge. The
forces are not drawn to scale; the forces composing the couple (a andb) would have to be much greater
than the retracting force (c) in order to produce clockwise rotation. d is the downward force component
explained in Fig. 11.

not sufficiently vertical to yield elevation, whereas that depicted in Fig. 13 shows the
necessary vertical direction but would, by itself, procline rather than retrocline.
Vertical anchorage from molars. Anchorage could be defined as that which experi-
ences and provides resistance to the reaction of a force system employed to achieve a
desired tooth movement. Horizontal movement (for example, incisor retraction) requires
horizontal force and horizontal anchorage. Vertical movement (for example, incisor in-
trusion) requires vertical force and vertical anchorage. Dental anchorage is rarely station-
ary. Molars may move mesially when used as anchorage for retraction of canines across a
premolar extraction space; nevertheless, they do provide anchorage. Molars provide
Volume 80
Number 5 Understanding, planning, and managing tooth movement 467

Fig. 13. A 50 Gm. intrusive force produces a clockwise moment of 850 Gm. mm.; a couple of 100 Gm.
forces produces a counterclockwise moment of 1,200 Gm. nm. The result is a combination of transla-
tion in the direction of the intrusive force and lingual tipping by the net 350 Gm. mm. counterclockwise
moment.

---- 12-13-79

Fig. 14. Tracings show the results of the first 6 months of treatment of a 15year-old girl. The movement
of the center of resistance of the maxillary incisor demonstrates the direction of the line of force which,
clearly, would have “flared’ the incisors, had it not been for the stronger opposite rotational effect of the
couple. An 0.016 inch round arch wire provided the intrusive force to the incisors.
Fig. 14A-D. A preformed mandibular “udder arch” stretched to fit the maxilla provided lingual crown
torque (A and B). Light vertical elastics connecting the posterior ends of the arch wires helped keep the
molars upright in resistance to the anchor bends. “Rat traps” can also provide the lingual crown torque
(C and D); here an 0.012 inch auxiliary is wound on an 0.016 inch arch wire.

“vertical anchorage” for incisor intrusion by their resistance to either the distal tipping
effects of anchor bends or the extrusive effects of elastics and anchor bends working in
combination. Occlusion probably augments molar resistance, especially in the latter case.
Understanding the principles of moments and couples enables us to gain insight into
the amount of force that is available for incisor intrusion by analyzing the interaction of
arch wires and elastics at the molars. Arch wires and molar tubes act as levers and fulcra.
Fig. 15 illustrates the forces that must be exerted by the dentoalveolar structures in
reaction to arch wires delivering 60 Gm. of intrusive force at the midline (that is, 30 Gm.
per side). In the maxilla, the 30 Gm. force acting at the end of the lever arm 36 mm. from
the fulcrum (mesial end of the molar tube) must be balanced by a 180 Gm. force in the
same direction at the distal end of the 6 mm. long tube (30 Gm. x 36 mm. = 180
Gm. x 6 mm.) and there must be a force of 210 Gm. in the opposite direction at the
mesial end of the tube, since the whole system is stationary (30 Gm. + 180 Gm. [down-
ward] = 210 Gm. [upward]). The lower arch is analyzed similarly. Now we can see that
an arch wire applying 60 Gm. of intrusive force to the incisors must subject each upper
molar to a couple producing a moment of 1,080 Gm. mm. and each lower molar to a
moment of 900 Gm. mm. plus an extrusive force on the mesial end of each tube of 30 Gm.
(which also adds a small increment to the moment).
In some instances, intermaxillary elastics act to counter the effects of the arch wires
upon the molars and thus reinforce vertical anchorage (for incisor intrusion). If a molar
does not tip mesiodistally, we can assume that the arch wire and elastic affecting it are
Volume 80
Number 5 Understanding, planning, and managing tooth movement 469

309 1809
v Y
3bmm A bmm
2109

1809

t-R --A
v
- 30mm bmmA
.
309 1509

Fig. 15. (See text.)


balanced in this plane and that they produce equal and opposite moments; thus, for a given
elastic, we can determine how much intrusive force an arch wire can deliver to the incisors
without tending to tip the molars distally. In practice, vertical anchorage is derived not
only from intermaxillary elastics as illustrated in Fig. 16 but also from the molars them-
selves in the following ways: (1) the molars may resist some force for some time (up to a
threshold of force x time) without significant movement; (2) some distal tipping of the
molars may be acceptable in some cases; (3) occlusion may impede molar tipping; (4)
since molars tend to tip mesially in Stages II and III, they may be tipped distally in Stage I
to a greater extent than would be acceptable in the final result.
For the following analyses, let
M,, = Moment produced by the elastic on the molar.
M, = Moment produced by lower arch wire on the molar.
M,, = Moment produced by the upper arch wire on the molar.
J? = Intrusive force that can be delivered to lower incisors if arch wire and elastics are in
balance (on each side of the arch).
Y = Intrusive force that can be delivered to upper incisors if arch wire and posterior
portions of the elastics are in balance (on each side of the arch); the effect of the
anterior end of the elastic is not considered here; it would decrease the effective
intrusive force of the arch wire upon the incisors.
Each elastic pulls with a force of 60 Gm. on the canine ring with teeth in occlusion.
Molar tube length = 6 mm. In Fig. 16, B and C, the resultant forces of the elastics on the
lower molars are calculated by vector analysis.
Fig. 16, A shows Class II elastics and molar tubes at ideal mesiogingival angulations:
M,, = 60 Gm. x 8 mm. = 480 Gm. mm.
M, =5X x6mm. +X x2mm. =32Xmm.
If the lower molar shows no tendency to tip mesially or distally, the moments pro-
duced by the arch wire and elastic must be of equal magnitudes in this plane. Thus:
B C
Fig. 16. (See text.)

M,, = 480 Gm. mm. = M-i = 32Xmm.


Therefore, X = 1.5Cm. (per side).

The lower arch wire can deliver 30 Gm. of intrusive force at the anterior midline
without tending to tip the molars distally. This figure, of course, applies only to situations
with precisely this geometry and elastic force magnitude.
The maxillary arch wire cannot deliver any intrusive force to the upper incisors
without tending to tip the upper molars distally, as there are no countermoments corre-
sponding to those produced by the Class II elastics at lower molars.
Volume 80
Understanding, planning, and managing tooth movement 471
Number 5

Fig. 17.A, Location of a moving “center of rotation, C and C’ as proposed by Hurd and Nikolai. B,
location of the true “center of rotation,” if such a concept has any validity, for the same tooth movement.
Note that C is not along the long axis.

Note that elastic placement is critical; if the elastic were hooked at the mesial portion
of the tube it would create a smaller moment and the intrusive force available from the
arch wire would be correspondingly less.
Fig. 16, B shows “check elastics,” ideally placed molar tubes, Class II, division 1.
“Check elastics,” so called because of their resemblance to the check mark (c/), have
one end hooked over the distal end of the upper arch wire, both strands under the distal
end of the lower arch wire, and the other end on the elastic hook mesial to the upper
canine. They will be discussed fully in the next paper in this series.
MANDIBLE.

Me,= 108 Gm. x 7.5 mm. = 810 Gm. mm.


M, =32Xmm. = 810 Gm. mm.
X = 25.3 Gm.
The lower arch wire can deliver 51 Gm. of intrusive force at the anterior midline.
The posterior segment of the elastic produces no significant moment on the
MAXILLA.
upper molar but tends to translate it downward and backward.
Fig. 16, C shows “check elastics,” ideally placed molar tubes, Class 1.
MANDIBLE

M,, = 87 Gm. x 6 mm. = 522 Gm. mm.


M, = 32Xmm. = 522Gm. mm.
X = 16.3 Gm.
The lower arch wire can deliver 33 Gm. of intrusive force at the anterior midline.
MAXILLA.

MeI = 60Gm. x 5 mm. = 300 Gm. mm.


Fig. 18. Continuous changes in force direction are necessary to maintain curvilinear motion of a body.

M,- = 6Y x 6 mm. + Y x 2 mm. = 38Y mm.


= 300 Gm. mm.
Y= 8Gm.
The upper arch wire can deliver 16 Gm. of intrusive force at the anterior midline;
however, note that only the effect of the posterior segment of the elastic on the maxilla has
been considered. The effect of the anterior segment is exactly like that of a Class II elastic,
tending to counteract the intrusive force of the arch wire.
Center of rotation. The use of the term center of rotation, generally without consid-
eration of its definition, is widespread in the orthodontic literature. It may be inappro-
priate, misleading, and confusing, making the study of biomechanics seem more compli-
cated than it need be. Christiansen and Burstone,lg Burstone,” Nikolai,x and Hurd and
Nikolai” have stated or implied that the center of rotation is always somewhere along the
line representing the long axis of the tooth. Burstone and his associates” recently stated
that this need not be so. Hurd and NikolaP have stated that it moves “along a path
coincident with a segment of a line in a position depicting the tooth angulation midway
through the movement. ’ ’
It may be true that in some cases an infinitesmal instantaneous movement effected by a
force perpendicular to the long axis of a tooth may center about a point somewhere along
the line representing the long axis of the tooth. However, there is no reason to assume and
no evidence to support the idea, that tooth movements of clinically significant magnitude
generally occur as rotations about axes located along this line. Nor is it logical or neces-
sary, in considering a discrete tooth movement, to think of a moving “center of rotation, ”
which is self-contradictory. If a point is moving, it is not the center of that movement.
Surely the “center” must be a fixed point about which the movement occurs.
Such a fixed “center of rotation” can be determined for the motion implied by any two
tooth positions through the application of basic geometry. The line segments connecting
corresponding points of the tooth in its two positions are chords of concentric circles about
the “center of rotation.” The perpendicular bisectors of these chords intersect at the
Understanding, planning, and managing tooth movement 473

Fig. 19A. (See text.)

i ii

C
Fig. 19B and C. (See text.)

center. Thus, a tooth movement for which Hurd and Nikolai would hypothesize a moving
“center of rotation” (Fig. 17,A) can be shown to have a fixed center (Fig. 17, B). In fact,
the tooth movement was not actually rotation about this center, but translation along a
straight line combined with rotation about its own CR, exactly the same movement shown
in Fig. 8. Any tooth movement can be “shown” to be a rotation about a center, but basic
physics implies that the movement was most likely linear with a superimposed rotation; it
is simplest to think of it as such.
“Center of rotation” implies curvilinear motion, which requires a force whose direc-
tion is constantly changing in a very specific manner. It is easiest to think of this force as
consisting of two components, one parallel to the direction of movement (tangent to the
curve) of the center of mass of the body (or CR, in tooth movement) and the other
ii

D
Fig. 19D and E. (See text.)

ii

F
G
Fig. 19F and G. (See text.)

perpendicular to it, toward the “center of rotation” (Fig. 18). This second component is
the centripetal force that holds the body to its curved path as gravitational attraction holds
the planets in their orbits about the sun.
The concept of “center of rotation” has never been proved. It has been assumed and
discussed widely, but no one has ever demonstrated that a macroscopic movement of a
tooth, produced by a constant force, has followed an arc of a circle. Nor has anyone ever
professed to explain what the physical basis for such a movement and the complex force
system that would be required to account for it might be.
There is no need to suppose that teeth generally move along curved paths. The simple
force systems applied in orthodontic treatment, as explained in the foregoing sections of
Vohmr 80
Understanding, planning, and manuging tooth movement 475
Number 5

J
Fig. 19H to J. (See text.)

this article, tend to effect translation along a straight line, rotation about the center of
resistance, or a combination of both in which the CR moves along a straight line and the
tooth rotates about CR. It is possible that the periodontal tissues may not respond uni-
formly or consistently, but there is no reason to assume that they dictate motion along
circular paths.

Appendix: Incisor movement exercises


Here we shall examine a number of different incisor positions to determine the best
force system for the correction of each. The anchorage requirements for some of these
cases may be such that extraoral traction would be indicated, but we will not consider that
aspect of their management here; we will limit our concern to force direction and type
rather than source. For simplicity, only the incisors are portrayed; their orientation in the
face is implied in the text for each case. The final position of the lower incisor represents
both movement of the incisor itself and molar extrusion.
Case A (Fig. 19, A) has “drooping” maxillary incisors at normal inclinations; they
require simple translation up and back along the bracket-CR line. Lower incisor position
relative to facial plane is to be maintained, and upper incisor movement is to account for all
bite opening. A simple force along the bracket-CR line would be ideal but difficult to
sustain enough to achieve this movement. Seeking this amount of overbite reduction
without any contribution from molar extrusion or lower incisor intrusion may be unrealistic.
Case B is similar to Case A, but less upward movement of the maxillary incisors is
needed because the molars and mandibular incisors contribute to bite opening (Fig. 19,
Bi). A force along the bracket-CR line would move the incisors up too much and back too
little; a force parallel to the desired CR movement would yield nearly pure tipping. What
is required is, first, a force passing just below CR producing excess intrusion and tipping
(Fig. 19, Bii) followed by lingual root torque (Fig. 19,Biii). The force causing the extrusion
that is an undesirable side effect of the torque is not illustrated.
For the proclined incisors in Case C, even though the molars and mandibular incisors
have contributed so much to bite opening that the maxillary incisal edges need only be
maintained at their original level, a force passingjust below CR is still required, but it may
be somewhat more horizontal than in Case B.
Case D differs from Case C in that more elevation of the maxillary incisors is required.
A more vertical force along the bracket-CR line would not effect the desired retroclina-
tion (Fig. 19, LX) but a more horizontal force or a couple (Fig. 19, Diii) lets the incisal
edges drop instead of giving the desired elevation. The ideal force system would combine
a force parallel to the desired CR movement with a lingual crown torque couple (Fig. 19.
Di).
Case E is much more severe, but projected movement of the mandibular incisors
makes the requirement for the maxillary incisors virtually the same as in Case D.
The incisor relation in Case F is identical to that of Case C and D but, as in Case A, it
would be best to accomplish all overbite reduction by elevation of the “droopy” maxillary
incisors. Again the force system should consist of a force parallel to the desired movement
of CR combined with a couple providing lingual crown torque. As in Case A, this
treatment goal may be unrealistic.
Case G is similar to Case E, but no protraction of the mandibular incisors and very
little retroclination of the maxillary incisors is permissible, so maxillary incisor movement
requires more horizontal translation. It could be treated as in Cases D and E, but this might
result in excessive elevation and a need for considerable lingual root torque. The alterna-
tive would be to treat it as shown (Fig. 19, G). Some lingual root torque (Fig. 19, Gii)
would probably be required after overjet correction, as in Case B. Either way, delicate
balance and copious anchorage would be required.
Case H shows what happens if the incisor relation portrayed in Cases C, D, and F is
treated with conventional Begg mechanics. Not only do the maxillary incisors not con-
tribute to bite opening but, because their edges drop, compensatory additional contribu-
tions are required of the mandibular incisors and molars. The upper incisors would fall
still further with the application of lingual root torque in Stage III.
Case I portrays the tooth movement that is supposed to occur in Begg Stage I. This is
unlikely with the force system commonly used. Such movement would require a nearly
vertical force (parallel to the proposed CR movement) plus a lingual crown torque couple,
as shown. In theory, at least, such a force system holds promise for use in cases in which
intrusion of proclined incisors is desired, and it is now being tested clinically (see Fig.
14).
Finally, in a Class II, Division 2 case such as Case J, overbite reduction is the primary
requirement. A nearly vertical force is required initially, to be followed by lingual root
torque after the bite is open. Active lingual root torque should wait until after the overbite
reduction, as all mechanisms for producing it tend to deepen the bite.

REFERENCES
1. Cadman, G. R.: A vade mecum for the Begg technique, AM. J. ORTHOD. 67: 477-512, 601-624, 1975.
2. Swain, B. F.: Begg differential light forces technique. In Graber, T. M., and Swain, B. F. (editors): Current
orthodontic concepts and techniques, ed. 2, Philadelphia, 1975, W. B. Saunders Company, vol. 2.
3. Williams, R.: The Begg technique. In Salzmann, J. A.: Orthodontics in daily practice, Philadelphia, 1974,
J. B. Lippincott Company.
4. Begg, P. R., and Kesling, P. C.: Begg orthodontic theory and technique, ed. 3, Philadelphia, 1977, W. 8.
Saunders Company.
l’olume 80
Number 5 Understanding, planning, and managing tooth movement 477

5. Begg, P. R., and Kesling, P. C.: The differential force method of orthodontic treatment, AM. J. ORTHOD.
71: l-39, 1977.
6. Yettram, A. L., Wright, K. W. J., and Houston, W. J. B.: Centre of rotation of a maxillary central incisor
under orthodontic loading, Br. J. Orthod. 4: 23-27, 1977.
7. Davidian, E. J.: Use of a computer model to study the force distribution on the root of the maxillary central
incisor, AM. J. ORTHOD. 59: 581-588, 1971.
8. Nikolai, R. J.: Periodontal ligament reaction and displacements of a maxillary central incisor subjected to
transverse crown loading, J. Biomech. 7: 93-99, 1974.
9. Baumrind, Sheldon: A reconsideration of the propriety of the “pressure-tension” hypothesis, AM. J.
ORTHOD. 55: 12-22, 1969.
10. Gianelly, A. A., and Goldman, H. M.: Biologic basic of orthodontics, Philadelphia, 1971, Lea & Febiger.
11. Burstone, Charles: Application of bio-engineering to clinical orthodontics. In Graber, T. M., and Swain, B.
F. (editors): Current orthodontic concepts and techniques, ed. 2, Philadelphia, 1975, W. B. Saunders
Company, vol. 1.
12. Weinstein, Sam: Minimal forces in tooth movement, AM. J. ORTHOD. 53: 881-903, 1967.
13. Hixon, E. H., et al.: Optimal force, differential force, and anchorage, AM. J. ORTHOD. 55: 437-457, 1969.
14. Storey, E., and Smith, R.: Force in orthodontics and its relation to tooth movement, Aust. J. Dent. 56:
11-18, 1952.
15. Hixon, E. H., et al.: On force and tooth movement, AM. J. ORTHOD. 57: 476-489, 1970.
16. Reitan, Kaare: Biomechanical principles and reactions. In Graber, T. M., and Swain, B. F. (editors):
Current orthodontic concepts and techniques, ed. 2, Philadelphia, 1975, W. B. Saunders Company.
17. Graber, T. M.: Dentofacial orthopedics. In Graber, T. M., and Swain, B. F. (editors): Current orthodontic
concepts and techniques, ed. 2, Philadelphia, 1975, W. B. Saunders Company.
18 Goldman, H. M., and Gianelly, A. A.: Histology of tooth movement, Dent. Clin. North Am. 16: 439-448,
1972.
19. Christiansen, R. L., and Burstone, C. J.: Centers of rotation within the periodontal space, AM. J. ORTHOD.
55: 353-369, 1969.
20. Hurd, J. J., and Nikolai, R. J.: Centers of rotation for combined vertical and transverse tooth movements,
AM. J. ORTHOD. 70: 551-558, 1976.
21. Burstone, C. J., Pryputniewicz, R. J., and Bowley, W. W.: Holographic measurement of tooth mobility in
three dimensions, J. Periodont. Res. 13: 283-294, 1978.

280 Great King St.

S-ar putea să vă placă și