You are on page 1of 22

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

,1 ,2 3
1
, , , ,
.
2
, , .
3
, , ,

Review of properties and clinical applications of orthodontic wires


Ilias M istakidis,1 Nikolaos Gkantidis,2 Nikolaos Topouzelis3
Dentist, Postgraduate Student in Dental Biomaterials, Department of Basic Dental Sciences, School of Dentistry, Aristotle University of
Thessaloniki, Greece.
2
Orthodontist, PhD candidate, University of Bern, Switzerland.
3
Associate Professor, Department of Orthodontics, School of Dentistry, Aristotle University of Thessaloniki, Greece.
1



.

.


, ,
.

,
. ''''
. , .



.

,

. ,
-, -, -
. ,
'''' -

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

ABSTRACT
During the last decades, a variety of alloys has been
used in orthodontics to manufacture wires. The orthodontic clinician is called to select from a large number
of materials that meet the biomechanical requirements of the clinical case to be treated.
Mechanical properties of orthodontic wires are
assessed by different laboratory tests, such as tensile,
torsional, and bending tests. Although wire characteristics determined by such tests cannot be directly
linked with their clinical application, they provide a
basis for useful comparisons. The "ideal" wire characteristics have been specified by a number of authors.
However, each wire may be considered ideal or not,
depending on the targeted clinical outcome on each
case. The clinician should know the properties and biomechanical behavior of available wires in order to
choose the appropriate wire depending on the targeted outcome in different orthodontic treatment phases.
The aim of this literature review is to summarize
orthodontic wire properties and demonstrate their
clinical applications as shown by their general properties. Stainless steel, cobalt-chromium, nickel-titanium,
beta-titanium and multistranded wires are studied.
Moreover, the so-called "aesthetic" wires are
reviewed, as well as their potential development
expected in the near future.
Key words: Orthodontic wires, mechanical properties,
applications
Hell Orthod Rev 2011;14:45-66.
Received: 31.03.2011 Accepted: 10.05.2011

45

/ Properties and applications of orthodontic wires

.
: O , ,
2011;14:45-66.
: 31.03.2011 : 10.05.2011

EIAH

.

.

, ,
(Houston
., 1996), /
0,15-5N (Shetty
., 1994; Holberg ., 2008). (Reitan, 1967; Van Leeuwen .,
2010).


(Leach ., 2001).


.


.

(Rudolph ., 2001). ,


.

, . 1930
.

46

HELLENIC ORTHODONTIC REVIEW

INTRODUCTION
During orthodontic treatment, the aim is to move the
teeth to a targeted position usually by applying forces to
them. An ideal force is the one that produces rapid tooth
movement without damage to the teeth or periodontal
tissues. Although it is difficult to precisely determine the
value of the ideal force since different biological and other
factors such as the type of movement and tooth size
should be considered (Houston et al., 1996), orthodontic/orthopedic forces usually range from 01.5-5N (Shetty
et al., 1994; Holberg et al., 2008). Studies have shown that
the application of lower forces produces the optimal
results. Application of excessive force exceeding vascular
blood pressure reduces cellular activity in periodontal tissues and slows down or stops tooth movement at least
for a period in time (Leach et al., 2001).
The ability to apply lower forces and achieve a wider
range of movements between sessions is of major importance to improve both the quality and performance of
treatment. The primary method to apply forces to the
teeth during orthodontic treatment using fixed appliances
is the orthodontic wire. The wire releases the energy
stored upon its placement by applying forces and torque
to the teeth through the appliances placed on them
(Rudolph et al., 2001). Therefore, knowledge of the biomechanical behavior and clinical applications of orthodontic wires is required to properly apply the treatment
plan.
The progress made in recent decades in orthodontic with
technology has resulted in a large variety of wires with a
wide range of properties. Up until 1930 all orthodontic
wires available were made of gold. Nowadays, these wires
are rarely used mostly due to their high cost compared to
that of alternatives. The use of stainless steel to produce
orthodontic wires was introduced in 1929. Stainless steel
quickly gained popularity over gold due to its improved
mechanical properties and lower cost. Since then several
other alloys exhibiting particular properties, as required,
have been introduced in clinical practice. These include
cobalt-chromium, nickel-titanium, beta-titanium and multistranded stainless steel wires.
No wire is appropriate for all treatment stages and no wire
is ideal. The clinician should consider a variety of wire
parameters and characteristics as necessary. The selection
of an appropriate wire size and type for each treatment
stage may, in theory, lead to a reduced number and duration of sessions, and possibly treatment as well, in cases
where the wire acts throughout the time period between
consecutive sessions, thus resulting in a higher-quality and
more predictable treatment outcome.
The properties of orthodontic wires are determined by different types of laboratory tests investigating wire behavior in terms of tension, bending, and torsion (Kapila and

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires


.

1929
. ,
,
.
-, -, -
.

.

.

,



.
,
, (Kapila Sachdeva, 1989).

,
.
1 2 .
, 3 .
, , ,
(Kusy Greenberg, 1981;
Asgharnia Brantley, 1986).

.

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

Sachdeva, 1989). Although these tests do not fully simulate the clinical setting where wires are used, they provide
a basis for comparison of wires. Bending tests provide
some information on wire behavior when they are subjected to 1st or 2nd order bends. Similarly, torsional test
results reflect, to a certain degree, wire behavior in 3rd
order bends. It should be noted that tension, bending, and
torsion are completely different stress states investigating
different characteristics related with wire performance,
respectively (Kusy and Greenberg, 1981; Asgharnia and
Brantley, 1986). Therefore, wire characteristics determined
under these three stress states should be assessed independently.
WIRE CHARACTERISTICS OF CLINICAL RELEVANCE
There are different characteristics considered desirable for
a wires good behavior depending on the clinical case,
such as aesthetics, resiliency, elasticity, a large spring-back,
low stiffness, high formability, high stored energy, biocompatibility, and stability in use, low friction, and the
capability to be welded with biomechanical attachments
(Kapila and Sachdeva, 1989; Kusy, 1997).
The maximum spring-back (range) is also referred to as
maximum elasticity, range of deflection or range of activation or working range, and it is the property related
with the ratio of yield strength to the modulus of elasticity of the material (YS/E). The yield strength is the stress
required to obtain 0.2% deflection of the wire after it is
unloaded. A large range of activation enables the clinician
to achieve larger activations resulting in the wires
increased working time (Ingram et al., 1986; Zufall and
Kusy, 2000a).
Stiffness or load deflection rate is the force magnitude
delivered by an appliance and is proportional to the modulus of elasticity (E) (Goldberg et al. 1983). Low wire stiffness provides the ability to apply lower forces, a more constant force overtime as the appliance experiences deactivation, and greater accuracy in applying a given force (Burstone and Goldberg, 1980).
Formability is the specific property allowing to easily form
the wire without breaking it or significantly altering its
characteristics.
Stored energy or resilience is the property of a material to
store energy while deflected and produce this energy
while unloaded. This feature represents the work available to move teeth.
Biocompatibility (or biohostability) includes resistance to
corrosion and tissue tolerance to wire elements. Ideally,
the wire should neither favor nor be a substrate for developing microorganisms and allergies.
The capability to be welded (or soldered) with biomechanical attachments is important in wires and may be
achieved through electric welding or by using a binding

47

HELLENIC ORTHODONTIC REVIEW

/ Properties and applications of orthodontic wires

, ,
, , (spring-back), , ,
,
,
(Kapila Sachdeva,
1989; Kusy, 1997).
(spring-back, range),
,
-,
(yield strength) (modulus of elasticity) (YS/E).

0,2%.


(Ingram ., 1986; Zufall Kusy,
2000a).

(stiffness or load deflection rate)

(E) (Goldberg
., 1983).
,

(Burstone Goldberg,
1980).
(formability)

.
(stored energy or resilience)


.
.
(biocompatibility, biohostability)
. ,
,
.
(welding or soldering)

48

agent.
Low friction is a major property when attempting to relatively move a tooth/bracket along a wire.
TYPES OF ORTHODONTIC WIRES

Stainless Steel (SS)Wires


Stainless steel alloy used to produce orthodontic materials
is AISI (American Iron and Steel Institute) type 302 or 304
(Khier et al., 1988). The stainless steel type commonly
used is also referred to as austenitic 18/8 suggesting its
chromium and nickel content of approximately 18% and
8%, respectively. This composition (mostly the chromium
content) allows the formation of a passivation oxide layer
blocking oxygen diffusion to the underlying mass, thus
making the alloy resistant to corrosion (i.e. biocompatibility) and ensuring a steady austenitic crystal structure for
steel (Brantley 2001).
Commercially available stainless steel wires have a range
of values both for elasticity and yield strength related to
the change in different parameters at their production
stage such as freezing and incandescence during cold
working.
The ratio of yield strength to the modulus of elasticity
(YS/E) indicates a lower spring-back of stainless steel wires
than those of beta-titanium or nickel-titanium. The stored
energy of activated stainless steel wires is also significantly lower than that of the other two types of wires
(Andreasen and Morrow, 1978; Drake et al., 1982). This
implies that stainless steel wires produce higher forces
applied during shorter time periods.
Stainless steel wires may be soldered with different biomechanical attachments, mostly through a binding agent,
although this process is highly demanding. New welding
techniques recommended for use in orthodontics without
a binding agent but with the use of Laser or TIG (tungsten
inert gas) welding produce satisfactory results, however,
they are currently too expensive and require the use of
sophisticated laboratory equipment (Bock et al., 2008).
As stated above, the corrosion resistance of stainless steel
is good in general. A low release of nickel and chromium
is observed, which, however is below the dietary average
intake (House et al, 2008). Nickel, mostly, but chromium as
well, may induce hypersensitivity reactions. There is an
apparently increasing interest in modern scientific literature with regard to orthodontic patients, although longterm experience from the use of nickel-containing orthodontic appliances does not justify the rather strong controversy over their safety and biocompatibility (liades
and Athanasiou, 2002; Mikulewicz and Chojnacka, 2010).
A recent meta-analysis concludes that orthodontic treatment is not related to an increased likelihood of hypersensitivity reactions to nickel unless there is a history of
skin piercing (Kolokitha et al., 2008).

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

.
(friction)
- .

Stainless Steel (SS)



AISI
(American Iron and Steel Institute) 302 304
(Khier ., 1988).
18/8, 18% 8%
, .
( ) (passivation oxide layer),
, ( ) (Brantley, 2001).



,

(cold working).
(YS/E)
- -.

(Andreasen Morrow,
1978; Drake ., 1982).
.

, ,

. ,
Laser
TIG ( ), ,

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

It has been found that stainless steel wires have a lower


bracket-wire friction than other types of wires (Kusy and
Whitley, 1990; Kapila et al., 1990; Krishnan and Kumar,
2004a), whereas experiments have achieved to further
reduce friction values using nanotechnology applications
(Redlich et al., 2008)
Australian Wires are a particular type of stainless steel
wires available in different grades representing gradually
increasing stored energy values (resiliency). These wires
are considered to be made of standard 18/8 stainless steel,
however a recent study (Pelsue et al., 2009) reports carbon
content up to 10 times higher than that in a standard
stainless steel orthodontic wire (up to 0.20%). This may
explain increased surface roughness, hardness, porosity,
and propensity for breakage during clinical bending, particularly for higher grades (Pelsue et al., 2009). This is why
higher grades, despite their high stored energy, are currently used solely as biomechanical attachments.
Recently, super stainless steels have been developed,
which have a lower nickel content, higher corrosion resistance, and improved mechanical properties. These desired
characteristics may widen the use of stainless steel wires
in the future since they are expected to supersede titanium wires in terms of high cost and low configuration ability (Oh et al., 2003).

Cobalt-chromium wires
Cobalt-chromium wires are commercially available in
grades based on their ability to store energy for a given
force. These wires have very good formability prior to heat
treatment, which takes place once they are configured,
thus increasing their stored energy and functionality
(Kusy, 1997). With only a few exceptions, non-heat treated cobalt-chromium wires have a smaller spring-back than
stainless steel wires of the same section (Ingram et al.,
1986). However, once subjected to heat treatment in a
dental oven (482 C for 7-12 minutes), this property is
improved (Ingram et al., 1986). Heating causes the accelerated hardening of the alloy, increasing its resistance to
deformation (Filmore and Tomlinson, 1979), and as a
result, the wire has mechanical properties similar to those
of stainless steel. High formability combined with
increased elasticity and yield strength following heat
treatment by 10% and 20-30%, respectively, has made
Blue Elgiloy, a cobalt-chromium wire type, popular in clinical practice.
Other properties of these wires show significant similarities to those of stainless steel wire properties (Meling and
Odegaard, 1998a; Kusy et al., 2001). Therefore, in most
cases, these wires may be replaced by stainless steel wires
of the same section, but of lower cost, which explains
their declining use in orthodontic clinical practice (Kusy
and Greenberg, 1981; Kusy et al., 2001). Moreover, with
regard to torque control, Meling and Odegaard (1998b)

49

/ Properties and applications of orthodontic wires

(Bock ., 2008).
,
.
,

(House ., 2008). ,
, .

,
(liades Athanasiou,
2002; Mikulewicz Chojnacka, 2010).
-

piercing (Kolokitha ., 2008).

-
(Kusy Whitley, 1990; Kapila
., 1990; Krishnan Kumar, 2004a),

(Redlich ., 2008)
Australian Wires
(resiliency). 18/8
, (Pelsue .,
2009) 10
(
0,20%). , ,

(Pelsue ., 2009).
' , , .
,
(super stainless steels) ,
.

,

50

HELLENIC ORTHODONTIC REVIEW

report that it is difficult to achieve with cobalt-chromium


wires since softer wires have a greater variability in their
dimensions depending on the manufacturer, and harder
ones have a high stiffness making the precise application
of torque to a specific part of the wire arch a highly
demanding process. Furthermore, caution is required
while welding biomechanic attachments since high temperature causes hardening and subsequent reduction in
the yield and tensile strengths (Filmore and Tomlinson,
1976). Therefore, it is recommended to use a low-fusion
binding agent (Kapila and Sachdeva, 1989).

Beta-titanium wires
The use of beta-titanium alloys as orthodontic wires has
been introduced in 1979 (Goldberg and Burstone, 1979;
Burstone and Goldberg, 1980). These wires are also
known as titanium-molybdenium alloy () (ORMCO,
Orange, CA, USA) or Titanium Niobium (ORMCO, Orange,
CA, USA), since they were the first ones to be manufactured and the only ones commercially available for several years.
These wires have a modulus of elasticity lower than half
of stainless steel wires and almost twice that of Nitinol
(Verstrynge et al., 2006; Juvvadi et al., 2010). A beta-titanium wire can be deflected almost twice as much as a similar stainless steel wire (Goldberg and Burstone, 1979; Burstone and Goldberg, 1980). Moreover, it delivers half the
amount of force applied compared to a stainless steel wire
of similar section (Goldberg and Burstone, 1979; Burstone
and Goldberg, 1980).
These wires demonstrate good formability, but should not
be strongly bent for there is a risk of breaking (Burstone
and Goldberg, 1980). Electrical welding of biomechanical
attachments is possible, but it should be made within a
specific voltage range (Nelson et al., 1987) using wide
electrodes (Donovan et al., 1984), so that wire properties
are not altered since overheating makes it brittle (Burstone and Goldberg, 1980; Nelson et al., 1987). According
to a recent study, beta-titanium wires are better in terms
of joinability than stainless steel wires since they demonstrate higher resilience and better surface and structural
characteristics, which indicates only a minor change in
wire properties after welding (Krishnan Kumar, 2004b).
Controversy over biocompatibility for beta-titanium wires
is reduced due to the absence of nickel, although their
resistance to corrosion is similar to that of cobalt-chromium and stainless steel wires. Their resistance to corrosion
is due to the formation of a surface passivation oxide layer
(Kim and Johnson, 1999), whereas it should be noted that
exposure to fluoride agents often used by orthodontic
patients leads to the degradation, subsequent corrosion,
and qualitative alteration of the wires surface (Walker et
al., 2007). Experimental studies have shown that this negatively affects the mechanical properties of deactivation of

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

(Oh ., 2003).

-
-
.
,
, (Kusy, 1997).
,
(Ingram ., 1986).
,
(482 C 7-12 ) (Ingram ., 1986).
, (Filmore Tomlinson, 1979)

. ,

10% 20-30% ,
-
lue Elgiloy .


(Meling Odegaard, 1998a;
Kusy ., 2001). ,

,
(Kusy
Greenberg, 1981; Kusy ., 2001). ,
(torque), Meling
Odegaard (1998b)
- , , ,
,
(torque) . ,
,

(Filmore Tomlinson, 1976). ,

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

these wires, as well as of those made of stainless steel and


NiTi (Kwon et al., 2005; Walker et al., 2005; Walker et al.,
2007). Meanwhile, their ion release is increased giving rise
to biocompatibility concerns. The duration of wire exposure to fluoride agents appears to play a major role
(Ogawa et al., 2004). However, most data about the effect
of fluoride agents on titanium materials comes from in
vitro studies with oversimplified assumptions of the simulation of oral cavity conditions. These findings should be
confirmed by in vivo studies (Fragou and Eliades, 2010).
Their high cost, as well as their bracket-wire friction,
which is higher than any other alloy, is considered a disadvantage (Kusy and Whitley, 1989; Kapila et al., 1990;
Cash et al., 2004; Krishnan and Kumar, 2004a), although
the development of beta-titanium wires appears to
improve the quality of surface finish, and subsequently,
friction values (Kusy et al., 2004).
A new wire made of an alpha-beta titanium alloy called
TiMolium (TP Labs) has been recently launched. According
to the study by Krishnan and Kumar (2004a), this wires
stiffness and other characteristics (such as elasticity and
yield strength) are between the values set for stainless
steel and beta-titanium wires.

Nickel-titanium or Ni-Ti wires


The first nickel-titanium wire (Nitinol) was introduced for
use in orthodontics by Andreasen and Hileman in 1971.
The first generation of NiTi alloys demonstrated the
"shape memory" effect, which refers to the wires property to return to a previously manufactured shape after
deflection when heated. This return occurs within a transitional temperature range (TTR) (Andreasen et al., 1985).
This effect has been described by Andreasen et al. (1985)
as "superelasticity". Superelasticity is the effect during
which the wire applies a low, constant force with a
"plateau" during its loading or unloading (ijima et al.,
2002). This definition was based on structural changes
occurring in the wire (phase transformation) during
changes in temperature or its loading process (Meling and
Odegaard, 2001).
In 1985, Burstone et al recommended the Chinese NiTi
(Ormco, Orange, CA, USA), which exhibited 4.4 times the
spring-back of stainless steel wires and 1.6 times the
spring-back of the original Nitinol and provided a constant
force in the middle of its deactivation range (Bradley et al.,
1996). One year later, Miura et al. (1986) introduced the
Japanese NiTi (Sentalloy, DENTSPLY GAC International,
Bohemia, NY) with similar properties (Brantley et al.,
1997). Finally, in the early 90s, GAC introduced a new
wire (NeoSentalloy) having a pure shape memory in oral
cavity temperature, whereas Ormco introduced CuNiTi at
that time (Gioka and Eliades, 2002).
With regard to the metallurgical structure of NiTi alloys,
there are two major phases and an intermediate one. The

51

/ Properties and applications of orthodontic wires


(Kapila Sachdeva, 1989).

-
- 1979 (Goldberg Burstone, 1979;
Burstone Goldberg, 1980).
titanium-molybdenium alloy
() (ORMCO, Orange, CA, USA) Titanium Niobium
(ORMCO, Orange, CA, USA),

.


o Nitinol (Verstrynge .,
2006; Juvvadi ., 2010). -

(Goldberg
Burstone, 1979; Burstone Goldberg, 1980). ,


(Goldberg Burstone, 1979; Burstone Goldberg,
1980).
,
(Burstone
Goldberg, 1980). , (Nelson ., 1987)
(Donovan ., 1984)
,
(Burstone
Goldberg, 1980; Nelson ., 1987).
, -

,

,

(Krishnan Kumar, 2004b).
- ,
- . H
(Kim
Johnson, 1999),
,
, -

52

HELLENIC ORTHODONTIC REVIEW

two major phases are austenitic NiTi (austenite) at high


temperatures and martensitic NiTi (martensite) at low
temperatures and high stress levels. The intermediate
phase, the R phase delays the transition from austenite to
martensite upon cooling until lower temperatures are
achieved (Gioka and Eliades, 2002).
Upon heating, As and Af temperatures are those at which
the transformation to austenite starts and finishes, respectively. Wire alloys with actual shape memory in vivo (such
as NeoSentalloy) have Af temperatures lower than oral
cavity temperatures, therefore their structure is always
austenitic during use. Non-superelastic (martensitic) wires
with no actual shape memory (such as Nitinol) have Af
temperatures much higher than 37 C, therefore they have
microstructures incompletely transformed to austenite at
oral cavity temperatures. Copper nickel-titanium wires
(Copper Ni-Ti) are available with three Af temperatures,
27, 35, and 40C. Therefore, the 27C wire is useful in
people with oral breathing, the 35C wire is activated at
normal body temperature and the 40C wire is only activated upon consumption of hot foods and beverages.
Finally, there are wires having a pseudo-elastic behavior,
during which their martensitic structure is transformed to
austensitic as a result of stress from the activation of the
wire. These wires (such as Nitinol SE) are superplastic, but
do not have a thermoelastic shape memory in vivo (Brantley, 2001).
Superelastic NiTi wires also have a pure shape memory
and are highly sensitive to changes in temperature affecting their mechanical properties (Wilkinson et al., 2002; De
Santis et al., 2008). The stress of these wires is increased
with heating and reduced with cooling (Meling and Odegaard 1998a; Airoldi et al., 1997; Sakima et al., 2006; Kusy
and Whitley, 2007). According to a study by Kusy and
Whitley (2007), the Orthonol (Rocky Mountain Orthodontics, Denver, CO, USA) and the three Cu-Ni-Tis demonstrate the stiffness of standard Nitinol at low temperatures developed in the oral cavity and the stiffness of TMA
at high temperatures. With regard to the three types of
Cu-Ni-Ti, their need to exist is questionable. A recent clinical trial failed to detect a significant difference in the correction of crowding between standard Ni-Ti and Cu-Ni-Ti
wires (Pandis et al., 2009). In the laboratory, the only
major difference is seen when placing the wire, where CuNi-Ti 40C applies a lower force and Cu-Ni-Ti 27 applies a
higher one (Wilkinson et al., 2002; Kusy and Whitley,
2007). Therefore, it would be wrong to consider that
forces applied by superelastic wires to teeth during their
use are constant. However, bone adaptation mechanisms
include the mechanical signal transfer achieved through
dynamic, not static loadings (Gross et al., 2002). This most
likely suggests that changing loadings of superelastic
wires induced by temperature changes may, after all, be
an advantage.

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

, (Walker .,
2007).

NiTi (Kwon ., 2005; Walker ., 2005;
Walker ., 2007). , .
(Ogawa .,
2004). ,


in vitro
.
in vivo (Fragou Eliades, 2010).

-
(Kusy Whitley, 1989;
Kapila ., 1990; Cash ., 2004; Krishnan
Kumar, 2004a), ,
(Kusy ., 2004).

- TiMolium (TP Labs).
Krishnan Kumar (2004a),

( , )
-.

- Ni-Ti
- (Nitinol) Andreasen
Hileman 1971. NiTi '' '',
.
(transitional
temperature range, TTR) (Andreasen ., 1985).

Andreasen . (1985) ''''.
, , ''plateau''
(ijima .,

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

The most important advantage of NiTi wires is their


increased elasticity allowing a wide deflection and activation range by delivering low forces (Andreasen and Morrow, 1978; Burstone et al., 1985). In general, nickel-titanium wires have a higher energy storage capacity than betatitanium or stainless steel wires when activated with the
same amount of bending or torque (Drake et al., 1982;
Brantley, 2001). Moreover, for a given amount of activation, they apply a lower and more constant force than
stainless steel or cobalt-chromium wires. However, the
levels of forces delivered vary significantly between NiTi
wires from different manufacturers (Nakano et al., 1999).
A study by Sakima et al. (2006) on 0.019 x 0.025-inch NiTi
wires found that the lowest and most constant levels of
force are delivered by Copper NiTi 40C and NeoSentalloy
200 gr. However, it was pointed out that these wires do
not deliver forces at a temperature below 35C, therefore,
they should not be used in subjects with oral breathing.
Low formability of these wires is a disadvantage. Any
desirable bending should be performed to such an extent
so as to achieve a permanent deflection rate. Moreover,
these wires are often fractured when bent over a sharp
edge (Andreasen and Morrow, 1978), whereas bending
also adversely affects the spring-back of wires (Lopez et
al., 1979). Furthermore, they are expensive and cannot be
welded or fused (Brantley, 2001). Nickel-titanium wires
develop an equally high bracket-wire friction to that of
beta-titanium wires, which is higher than that of stainless
steel or chromium-cobalt wires (Kapila et al., 1990).
NiTi wires have a passivation oxide layer (titanium oxide),
which is more stable than that formed on steel wires
(chromium oxide) and provides them good corrosion resistance in the mildly acidic oral environment. However, it
was found that a low release of Ni or Ti ions may be seen
(Huang et al., 2003). According to a study by Eliades et al.
(2000), the intra-oral exposure of NiTi wires alters the
topography and structure of their surface. However, it has
not been fully clarified whether this may be of clinical relevance, possibly affecting the wires superelastic properties (Gioka and Eliades, 2002). It has been found that
(stress) increases the corrosion rate of NiTi and beta-titanium wires and that stress/deflection changes related to the
phase transformation of superelastic NiTi wires may affect
their corrosion rate in a way different than that of wires
not subject to phase transformation (Segal et al., 2009).
Additionally, the longer nickel-titanium wires remain in
the oral cavity, the more likely they are to fail due to normal wear, whereas it seems that larger square- or rectangular-section wires are shown to be more vulnerable compared to smaller-section wires (Bourauel et al., 2008).
Several clinical trials have shown that in many cases superelastic wires are not capable of demonstrating their superelastic properties in vivo, therefore, they do not have significant advantages over standard wires NiTi (Meling and

53

/ Properties and applications of orthodontic wires

2002).
(
) (Meling
Odegaard, 2001).
1985 Burstone . NiTi (Ormco, Orange, CA, USA) 4,4
1,6
Nitinol
(Bradley
., 1996).
Miura . (1986), NiTi (Sentalloy
DENTSPLY GAC International, Bohemia, NY) (Brantley ., 1997). ,
90 GAC
(NeoSentalloy)

CuNiTi Ormco (Gioka Eliades,
2002).

NiTi,
. NiTi ()
NiTi () .
, R,
(Gioka Eliades,
2002).
, As Af
. in vivo ( NeoSentalloy)
Af
. (),
( Nitinol), Af 37 C
. - (Copper Ni-Ti) Af, 27, 35 40C. ,
27C ,
35C 40C
.
, ,

54

HELLENIC ORTHODONTIC REVIEW

Odegaard, 1998a,b; Brantley, 2001).

Multistranded wires
Multistranded wires are made of a varying number of
stainless steel wire strands coaxially placed or coiled
around each other in different configurations. The stiffness of a triple-stranded 0.0175-inch wire is comparable to
that of a single-stranded 0.010-inch stainless steel wire,
whereas the triple-stranded is more resilient by 25%.
Moreover, a triple-stranded 0.0175-inch wire has similar
stiffness to that of a 0.016-inch nickel-titanium wire,
although the nickel-titanium wire can accept more than
50% greater activation than the multistranded one. Furthermore, the triple-stranded wire is 50% less stiff than a
0.016-inch beta-titanium wire (Kusy and Dilley, 1984). The
elastic behavior of coiled multistranded wires resembles
more the behavior of a single strand than of a set of
strands (Rucker and Kusy, 2002a,b,d). Multistranded wires,
such as nickel-titanium wires, have a spring-back, which is
rather independent of the wires size, as opposed to multistranded stainless steel wires (Ingram et al., 1986; Kusy
and Dilley, 1984; Kusy, 1981).
Taneja et al. (2003), following a laboratory study of different multistranded stainless steel wires concluded that
although the arrangement, section, and number of
strands should in theory greatly affect the range of forces
applied in a given deflection (Rucker and Kusy, 2002d), in
practice, this only occurs in deflections over 2 mm. Additionally, these investigators observed that coiled wires
deliver higher forces than coaxial wires.
Low forces developed by multistranded wires, low stiffness and a high spring-back are characteristics similar to
those of the more expensive titanium alloys (Wilkinson et
al., 2002; Rucker Kusy, 2002a), and could be an economically advantageous alternative (Kusy and Dilley,
1984; Kusy and Stevens, 1987). However, multistranded
wires do not always deliver low and rather constant forces
when unloaded, particularly after significant deflection
(Taneja et al., 2003). With regard to friction, multistranded stainless steel wires develop higher bracket-wire friction than NiTi wires and approximately more than 30%
greater friction than single-stranded stainless steel wires
(Rucker and Kusy, 2002c).
CLINICAL APPLICATIONS OF ORTHODONTIC WIRES
The proper selection of the type and size of wire can maximize efficiency of the clinical applications of orthodontic
wires in a way that all clinical requirements are fulfilled.
The wire sequence that is followed in the course of treatment is depended on the treatment technique (i.e., segmental or straight wire technique) and can be accordingly
modified. Knowledge of basic wire properties and biomechanical aspects can help the orthodontist select the wire

HELLENIC ORTHODONTIC REVIEW 2010 VOLUME 13 ISSUE 1 & 2

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires


.
( Nitinol SE) , in vivo
(Brantley, 2001).
NiTi
, (Wilkinson ., 2002; De Santis
., 2008).
(Meling
Odegaard 1998a; Airoldi ., 1997; Sakima
., 2006; Kusy Whitley, 2007). Kusy Whitley (2007), Orthonol (Rocky
Mountain Orthodontics, Denver, CO, USA)
Cu-Ni-Ti Nitinol
TMA
. Cu-Ni-Ti,
.

Ni-Ti
Cu-Ni-Ti (Pandis ., 2009). ,
, Cu-Ni-Ti 40C
Cu-Ni-Ti 27 (Wilkinson ., 2002; Kusy Whitley,
2007). ,
. ,
(Gross
., 2002).

.
NiTi

, (Andreasen Morrow, 1978;
Burstone ., 1985). ,

(Drake ., 1982;
Brantley, 2001). , ,
-

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

material, geometry and size that is optimal for each case


or make various technique combinations that will optimize the quality of the treatment provided. Usually, in the
initial stages of orthodontic treatment the most common
wires used are superelastic NiTi wires that are replaced by
beta- titanium and stainless steel wires later in treatment
(Keim et al., 2002). The high modulus of elasticity of Co-Cr
and stainless steel wires implies that this type of wires
exert twice the force of beta- titanium and 4 times the
force of the conventional NiTi wires for the same amount
of activation (Drake et al., 1982).
Studies on the elastic properties of the wires specify that
their mechanical behavior is determined by the following
equation: strength = stiffness x range (Kusy and Greenberg, 1981). For all the wires that follow linear elasticity
these characteristics are specified as following: (1)
strength: the maximum force that can cause elastic deformation of the wire, (2) stiffness: the strength to range
ratio, that is equal to force per activation unit and (3)
range: the maximum distance that a wire can be elastically activated. For all wires, except for some specific types of
NiTi wires, the relationship between force and elastic
deformation is linear (Zufall and Kusy. 2000a). It must be
emphasized that all three elastic properties of the wires
(strength, stiffness, range) depend on changes of the
geometry, the size and the length of the wire. For the
majority of the conventional wires, flexibility and working
range are decreased by increases in size, in contrast to
stiffness that is increased. This is not the always the case
for multistranded stainless steel wires and especially for
superelastic NiTi wires (Ingram et al., 1986). For all other
wires changes that are related to size and shape are of the
same degree irrespective of the construction material. For
example, half reducing the diameter of a stainless steel
will accordingly reduce its stiffness to a specific percentage of the initial stiffness. The same will happen for a
beta- titanium wire: half reducing its diameter will give
the same percentile reduction of its initial stiffness as for
the stainless steel wire (Proffit et al., 2007). In addition,
increasing the length of the wire will proportionally
decrease the stiffness but exponentially increase the elasticity and working range. Finally resilience is significantly
increased when the wire is loosely ligated into the brackets allowing some sliding freedom ( Adams et al., 1987).
It must be emphasized that rounding the edges of rectangular wires seriously reduces torque control and stiffness.
Torque control depends on multiple factors that are related to the bracket and its design, the type of ligation, the
free-space (play) between the wire and the bracket slot.
The raw material, the wire section and the edge bevel
appear to play a significant role (Sebanc et al., 1984; Gioka
and Eliades, 2004; Morina et al., 2008). In small activations
(<12) no difference was noted in torque expression
between stainless steel, TMA and Cu-Ni-Ti wires, whether

55

/ Properties and applications of orthodontic wires

. ,

NiTi (Nakano
., 1999). Sakima . (2006)
NiTi 0,019 x 0,025 ,
Copper NiTi 40C
NeoSentalloy 200 gr. ,

35C '
.

.

. ,

(Andreasen Morrow, 1978),

(Lopez ., 1979). , (Brantley, 2001).
-

-,
(Kapila ., 1990).
NiTi ( )

( )

. ,

Ni Ti (Huang ., 2003).
Eliades . (2000),
NiTi
. , ,

(Gioka Eliades, 2002). (stress)
NiTi - /

iTi
(Segal ., 2009). ,
- , -

56

HELLENIC ORTHODONTIC REVIEW

in higher angles (>24) stainless steel wires express 1.5-2


times and 2.5-3 times higher forces than TMA and Cu-NiTi respectively (Archambault et al., 2010). It was found
that superelasticity expression of NiTi wires at torsion in
mouth temperature starts at 20 degrees, even though
large variations appear depending on the manufacturer.
This evidence, in conjunction with the free play of the wire
inside the bracket slot, makes the clinical determination of
the relationship between the NiTi wire cross section and
the real torsion force extremely difficult (Partowi et al.,
2010).
According to Burstone (1981), the desirable stepwise
increase in wire stiffness in the treatment progress can be
attainted not only by proper wire configuration and size
increase, but by the selection of higher modulus of elasticity wires as well (Variable- modulus orthodontics). Various laboratory studies have shown that the expressed
force and stiffness of the wire vary proportionally (Rucker
and Kusy, 2002a, b; Garrec et al., 2005). The stiffness correlation of Cr-Co, SS, beta-Ti and NiTi wires is
1.2:1:0.42:0.26. The stiffness of multistranded wires is
1/25-1/5 the stiffness of a single strand stainless steel wire
(Burtsone, 1981). This allows the clinician to use bigger,
even square or rectangular wires, in earlier treatment
stages giving better three dimensional control of the
movements. Furthermore, according to Kusy (1997), keeping relatively constant the angular relationship between
the wire and the bracket slot helps in controlling the
amount of friction. However, even though low stiffness
and wide working range are desirable properties during
tooth movement, the wire action in areas where no movement is expected, as in the final stages of treatment, is
better controlled with wires that have narrower activation
range and higher stiffness (Burstone, 1985).
During the initial stages of treatment wires with wide
range of activation and low force delivery are preferred in
order to level and align the teeth (Kusy, 1997). Typically
the clinician selects a wire of low stiffness, high resilience
and satisfactory strength (Rucker and Kusy, 2002b). The
first wire to be selected should express the maximum
force that can be tolerated once attached at the pointbracket of maximum activation (Johnson, 2003). Theoretically, the initial wires should be round superelastic NiTi
wires of small diameter, that can later on be replaced by
rectangular NiTi wires which are relatively stiffer. NiTi
wires have high resilience that allows a wide range of
elastic deformation and activation (Andreasen anf Morrow, 1978; Drake et al., 1982; Brantley, 2001) and they
apply lower and more constant force in comparison to
stainless steel or cobalt-chromium wires for a given
amount of activation. According to Andreasen and Morrow (1987), the clinical advantages of NiTi wires include
less wire changes, less chair time, faster leveling and rotation correction and higher comfort for the patient. It is

HELLENIC ORTHODONTIC REVIEW 2010 VOLUME 13 ISSUE 1 & 2

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires


(Bourauel ., 2008).


in vivo

NiTi (Meling Odegaard,
1998a,b; Brantley, 2001).




. 0,0175

0,010 ,
25% . ,
0,0175
- 0,016
- 50%
. ,
50% - 0,016 (Kusy Dilley, 1984).
(Rucker Kusy,
2002a,b,d). , -,
, (Ingram ., 1986; Kusy Dilley,
1984; Kusy, 1981).
Taneja . (2003),
, ,
,
(Rucker Kusy, 2002d), 2 ..
,
.

,
(Wilkinson
., 2002; Rucker Kusy, 2002a)

EHNIKH OPOONTIKH EIEPHH 2010 TOMO 13 TEYXO 1 & 2

suggested that the use of rectangular wires that fit firmly


in the slots of the brackets should be avoided during the
initial leveling and alignment of the teeth because it
results in undesirable "reciprocal" movements of the
roots, thus increasing treatment time and root resorption
risk. During this phase it is preferred that the crowns of
the teeth are first arranged with the use of low diameter
round wires (Proffit et al., 2007). Then, rectangular NiTi
wires of higher cross section can be used, in early stages
of treatment, that will allow leveling, alignment, rotation
correction and torque control at the same time.
In particular cases, intermediate sizes of Australian wires
can be used in this phase for the correction of incisors
overbite (utility arches or reverse curve of Spee arches),
especially because of their high resistance to plastic deformation from external factors, like mastication forces
(Drake et al., 1982). The high modulus of elasticity and
stiffness of these wires dictate that wires of smaller diameter or of increased length, by means of forming loops,
should be used for the alignment and leveling of the dental arches (Krishnan and Kumar, 2004a; Verstrynge et al.,
2006).
Whatever wire is selected for the initial alignment of
teeth, it should be loosely ligated so that the wire is free
to slide inside the bracket slots. In that way this particular
stage of treatment appears to be significantly faster.
Superelastic NiTi wires and wires that express shape memory are useful in cases where large amounts of deformation are needed in order to insert the wire in the brackets,
as in high crowding cases (Bartzela et al., 2007). It should
be noted that the use of less stiff wires, like superelastic
NiTi wires, in the initial stages of treatment is suggested in
cases where the existing crowding is relatively symmetric.
Otherwise, there is a risk that alignment of the teeth will
at the same time lead to loss of dental arch shape (Proffit
et al., 2007).
Kusy and Stevens (1987) mention that a three stranded
0.015 stainless steel wire has a wider range of action
than conventional NiTi and beta-titanium wires of the
same or larger dimensions. The same authors state that
multistranded wires can be quite effectively compared to
NiTi wires and are a possible good alternative to more
expensive wires for the initial alignment (Quintao et al.,
2009). Clinical studies comparing superelastic NiTi wires,
conventional NiTi wires and multistranded stainless steel
wires did not find any significant difference regarding the
ability and speed of initial alignment and leveling of teeth
(West et al., 1995; Evans et al., 1998). Furthermore, two
systematic reviews on alignment efficiency and pain triggering comparing several types of single strand NiTi wires
and stainless steel wires did not reveal any significant differences (Riley and Bean, 2009; Wang et al., 2010).
Intermediate stages of treatment often demand proper
wire bending and bonding of various biomechanical acces-

57

/ Properties and applications of orthodontic wires

(Kusy Dilley,
1984; Kusy Stevens, 1987). ,
,
(Taneja .,
2003). ,

-
NiTi 1/3 (Rucker Kusy,
2002c).



.


(..
straight-wire ). , ,
,

. , NiTi, - (Keim
., 2002). -
2
- 4


(Drake ., 1982).

: strength = stiffness x range
( = x ) (Kusy Greenberg,
1981). (1) strength:
, (2) stiffness: strength
range, (3) range:
. NiTi, -

58

HELLENIC ORTHODONTIC REVIEW

sories. Beta- titanium wires, despite their high cost, are


characterized by very good forming behavior which, in
addition to their weldability without the need of a solder,
makes them wires of choice whenever a construction of a
complex configuration is needed. Furthermore, they exert
a wider range of activation and lower forces than stainless
steel or Co-Cr wires (Kapila Sachdeva, 1989; Kusy,
1997; Johnson, 2003). In cases where no wire bending is
required NiTi wires of large cross section can be inserted
(Kusy, 1997). Whenever sliding is desirable low friction
helps the movement. In this case smaller round stainless
steel wires are considered more appropriate (Kapila
Sachdeva, 1989; Kusy, 1997).
At this point it should be noted that the resistance in the
movement of a bracket along the wire during treatment
only partly is a result of friction (Burrow, 2009). Laboratory studies show that the binding of the wire that happens
shortly after the onset of tooth movement plays an important role (Thorstenson and Kusy, 2002; Hamdan and Rock,
2008), and the permanent deformation caused by the
notching of the wire, that temporarily stops the movement, is also important (Articolo ., 2000; Burrow,
2009). As a result, the resistance to tooth movement is not
a simple phenomenon that can be easily simulated and
controlled on laboratory tests of static or kinetic friction.
On the contrary this is a multifactorial phenomenon of
binding and release that is difficult to be successfully simulated in vitro in order to be sufficiently investigated
(Swartz, 2007; Burrow, 2009). As a result, all studies that
show stainless steel wires to be superior to all other wires
especially Ti-alloy wires, as far as friction is concerned, are
probably of small clinical importance (Peterson .,
1982; Kusy Whitley, 1990; Kusy Whitley, 1989;
Kapila ., 1990; Cash ., 2004; Krishnan
Kumar, 2004a).
In the final stages of treatment, where smaller movements are needed, the greater stability and torque control, formability and stiffness of stainless steel (or Cr-Co)
wires make them wires of choice (Kapila and Sachdeva,
1989; Kusy, 1997). In cases where the corrections that
must be done are larger, it is better to use beta-titanium
wires in order to have greater range of activation and to
avoid putting excessive forces on the teeth (Johnson,
2003). The low stiffness of NiTi wires makes them unsuitable for the final stages of treatment.
Sometimes, rectangular multistranded stainless steel
wires are used for the final settling of dental arches
because they allow small dental movements while keeping the final position of teeth.
Furthermore, multistranded wires are frequently used for
retention purposes after the orthodontic treatment,
where usually round wires (0.0215) bonded on the front
teeth are used (Zachrisson, 2007). It is suggested that the
wire should be submitted to some kind of thermal work-

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires


(Zufall Kusy, 2000a).
(strength, stiffness, range)
(),
.

,
.

NiTi
(Ingram ., 1986). ,


. , 50%,

. , - 50%,
(Proffit ., 2007). ,

, ,

. ,
, (Adams .,
1987).
, ,
(torque) . O (torque)

, ,
-
. ,
(edge
bevel) (Sebanc
., 1984; Gioka Eliades 2004; Morina .,
2008). (<12),
(torque)
, Cu-Ni-Ti,
(>24)
1,5-2
2,5-3
Cu-Ni-Ti (Archambault ., 2010).
i-Ti
20 ,

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

ing before its final fitting on the teeth in order to remove


any residual stresses.
ADVANCES AESTHETIC WIRES
Up until now, an aesthetic wire accompanied with the
required mechanical properties making it essentially useful for progress in orthodontics has not been found. The
two types of aesthetic wires available are metal-coated
and fiber-reinforced wires. Of these two, those available
for clinical use are the metal-coated ones, since fiber-reinforced ones are still in experimental stage. The materials
used for coating are synthetic fluoride resins, Teflon, or
epoxy polytetrafluoroethylene resins (Elayyan et al., 2008;
Elayyan et al., 2010). The so-called "white wires" are
prone to chewing forces or the enzymatic activity of the
oral cavity within three weeks of use in vivo (Kusy, 1997).
Color instability of these wires and exposure of the underlying metal is also often reported (Lim et al., 1994). It has
been found that 25% of coating is lost in 33 days intraorally, therefore, the wire becomes aesthetically degraded
(Elayyan et al., 2008). The coating seems to initially lower
bracket-wire friction (Husmann et al., 2002), however, surface roughness values soon increase in clinical use
(Elayyan et al., 2008). Coated wires deliver lower forces
when loaded and unloaded compared to non-coated
wires of the same nominal diameter (Elayyan et al., 2010).
This finding, even though expected (due to coating thickness, a slightly lower section wire is used), should be
taken into account since a large number of orthodontics
use wire diameter as their guide in certain clinical cases
(Elayyan et al., 2010).
The use of synthetic materials including matrixes reinforced with glass or others fibers is much promising (Zufall
and Kusy, 2000b; Goldberg and Burstone, 1992). The stiffness of these wires may be determined when manufactured without changing the wires dimensions, by changing fiber geometry and/or the ratios of fiber and matrix
materials and properly polymerising them (Imai et al.,
1998; Zufall and Kusy, 2000b). This makes it possible to
manufacture even rectangular aesthetic wires (Fallis and
Kusy., 2000). A recent study of a 3-point bending test on
fiber-reinforced and stainless steel wires led to encouraging results (Cacciafesta et al., 2008).
Since synthetic materials replace metal alloys in aeronautics, something similar is expected to occur to a certain
extent in orthodontics in the near future (Kusy, 1997).
Recently, in the wider scientific field of materials, polymers
with shape memory have been produced and used in a
variety of biomedical applications (Lendlein et al., 2005).
These materials may be a viable alternative to orthodontic wires currently used in the future (Eliades, 2007). A
novel (experimental) type of aesthetic materials for orthodontic use referred to as self-reinforced polymer compos-

59

/ Properties and applications of orthodontic wires

. ,

i-Ti

(Partowi .,
2010).
Burstone (1981),



,
(Variablemodulus orthodontics).
(Rucker Kusy,
2002a, b; Garrec ., 2005).
-, , - -
1,2:1:0,42:0,26.
1/25-1/5 (Burstone,
1981).
,
,
. , Kusy (1997), -
.
,

,
,
,

(Burstone, 1985).



(Kusy, 1997). , , (Rucker Kusy,
2002b).
-

(Johnson, 2003). ,
NiTi , , NiTi ,

60

HELLENIC ORTHODONTIC REVIEW

ites is also of particular interest. They are made of


polyphenylene, which demonstrates time-dependent
relaxation. Subsequently, their potential use in clinical
practice will require the expert orthodontist to understand
this behavior of wires (Goldberg et al., 2010).
CONCLUSIONS
Currently, a large variety of alloys is used to produce
orthodontic wires. Each type of wire demonstrates certain
properties, which sometimes vary significantly between
manufacturers. Different alloy properties combined with
the variety in sizes and geometries make the selection of
the proper wire for a specific clinical case a highly
demanding process.
Knowledge of the mechanical and physical properties of
wires, which, to some extent, determine their clinical
behavior is required to achieve a satisfactory and predictable outcome. Usually, as treatment progresses, wires
with gradually increasing stiffness are chosen. However,
since the ideal wire has not yet been discovered, each wire
may be an appropriate choice or not, depending on the
treatment stage and the desired clinical outcome. The
selection and sequence of wires should neither be made
irrationally nor be determined in advance for all patients;
instead, it should rely on current scientific evidence and
the clinical requirements of each case.
Advances in the field of biomaterials have provided
encouraging results with regard to the creation of aesthetic wires with satisfactory properties. It seems that in
the near future these wires will be available for orthodontics, thus increasing the range of available options.
References
Adams DM, Powers JM, Asgar K. Effects of brackets and ties on stiffness of an arch wire. Am J Orthod Dentofacial Orthop
1987;91:131-6.
Airoldi G, Riva G, Vanelli M, Filippi V, Garattini G. Oral environment
temperature changes induced by cold/hot liquid intake. Am J
Orthod Dentofacial Orthop 1997;112:58-63.
Andreasen GF, Hileman TB. An evaluation of 55-cobalt substituted
wire for orthodontics. J Am Dent Assoc 1971;82:1373-5.
Andreasen G, Hileman H, Krell D. Stiffness changes in thermodynamic nitinol with increasing temperature. Angle Orthod
1985;55:120-6.
Andreasen GF, Morrow RE. Laboratory and clinical analyses of nitinol
wire. Am J Orthod 1978;73:142-51.
Archambault A, Major TW, Carey JP, Heo G, Badawi H, Major PW. A
comparison of torque expression between stainless steel, titanium molybdenum alloy, and copper nickel titanium wires in
metallic self-ligating brackets. Angle Orthod 2010;80:884-9.
Articolo LC, Kusy K, Saunders CR, Kusy RP. Influence of ceramic and
stainless steel brackets on the notching of archwires during clinical treatment. Eur J Orthod 2000;22:409-25.

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

. Ni-Ti

(Andreasen Morrow, 1978; Drake .,
1982; Brantley, 2001),
-. ,


. Andreasen Morrow
(1978) - , , .

'''' ,
.

,
(Proffit ., 2007). ,
Ni-Ti , , , ,
(torque).
, ( utility
Spee),
, (Drake ., 1982).



(Krishnan
Kumar, 2004a; Verstrynge ., 2006).
, ,
.
.

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

Asgharnia MK, Brantley WA. Comparison of bending and tension


tests for orthodontic wires. Am J Orthod Dentofacial Orthop
1986;89:228-35.
Bartzela TN, Senn C, Wichelhaus A. Load-deflection characteristics of
superelastic nickel-titanium wires. Angle Orthod 2007;77:991-8.
Bock JJ, Fraenzel W, Bailly J, Gernhardt CR, Fuhrmann RA. Influence
of different brazing and welding methods on tensile strength
and microhardness of orthodontic stainless steel wire. Eur J
Orthod 2008;30:396-400.
Bourauel C, Scharold W, Jger A, Eliades T. Fatigue failure of asreceived and retrieved NiTi orthodontic archwires. Dent Mater
2008;24:1095-101.
Bradley TG. A differential scanning calorimetrc determination of the
phase transformation temperature ranges in superelastic and
nonsuperelastic nickel-titanium orthodontis arch wires. Master
of science Thesis. Columbus Ohio: Columbus University, 1993.
Bradley TG, Brantley WA, Culbertson B. Differential scanning
calorimetry (DSC) analyses of superelastic and nonsuperelasticity nickel-titanium orthodontic wires. Am J Orthod Dentofacial
Orthop 1996;109:589-97.
Brantley WA. Orthodontic wires, in: Brantley WA, Eliades T, eds.
Orthodontic materials: scientific and clinical aspects. Stuttgard:
Thieme, 2001: 91-9.
Brantley WA, Webb CS, Soto U, Cai Z, McCoy BP. X-ray diffraction
analyses of Copper Ni-Ti orthodontic wires. J Dent Res
1997;76:401.
Burrow SJ. Friction and resistance to sliding in orthodontics: a critical
review. Am J Orthod Dentofacial Orthop 2009;135:442-7.
Burstone CJ, Goldberg AJ. Beta titanium: a new orthodontic alloy.
Am J Orthod 1980;77:121-32.
Burstone CJ, Qin B, Morton JY. Chinese NiTi wire: A new orthodontic
alloy. Am J Orthod 1985;87:445-52.
Burstone CJ. Variable-modulus orthodontics. Am J Orthod
1981;80:81-6.
Cacciafesta V, Sfondrini MF, Lena A, Scribante A, Vallittu PK, Lassila
LV. Force levels of fiber-reinforced composites and orthodontic
stainless steel wires: a 3-point bending test. Am J Orthod Dentofacial Orthop 2008;133:410-3.
Cash A, Curtis R, Garrigia-Majo D, McDonald F. A comparative study
of the static and kinetic frictional resistance of titanium molybdenum alloy archwires in stainless steel brackets. Eur J Orthod
2004;26:105-11.
De Santis R, Dolci F, Laino A, Martina R, Ambrosio L, Nicolais L. The
Eulerian buckling test for orthodontic wires. Eur J Orthod
2008;30:190-8.
Drake SR, Wayne DM, Powers JM, Asgar K. Mechanical properties of
orthodontic wires in tension, bending and torsion. Am J Orthod
1982;82:206-10.
Elayyan F, Silikas N, Bearn D. Ex vivo surface and mechanical properties of coated orthodontic archwires. Eur J Orthod 2008;30:6617.
Elayyan F, Silikas N, Bearn D. Mechanical properties of coated superelastic archwires in conventional and self-ligating orthodontic
brackets. Am J Orthod Dentofacial Orthop 2010;137:213-7.
Eliades T, Athanasiou AE. In vivo aging of orthodontic alloys: implications for corrosion potential, nickel release, and biocompatibility. Angle Orthod 2002;72:222-37.
Eliades T, Eliades G, Athanasiou A, Bradley TG. Surface characterization of retrieved NiTi orthodontic archwires. Eur J Orthod

61

/ Properties and applications of orthodontic wires

-
, (Bartzela ., 2007).
,
NiTi,
, . ,

(Proffit .,
2007).
usy Stevens (1987) 0,015 - .
-

(Quintao
. 2009). , NiTi, NiTi ,


(West ., 1995; Evans ., 1998).
,

-
(Riley
Bearn, 2009; Wang ., 2010).


.
-
,
. ,

- (Kapila Sachdeva, 1989; Kusy,
1997; Johnson, 2003).

- (Kusy,
1997).
-

62

HELLENIC ORTHODONTIC REVIEW

2000;22:317-26.
Eliades T. Orthodontic materials research and applications: part 2.
Current status and projected future developments in materials
and biocompatibility. Am J Orthod Dentofacial Orthop
2007;131:253-62.
Evans TJW, Jones ML, Newcombe RG. Clinical comparison and performance perspective of three aligning arch wires. Am J Orthod
Dentofacial Orthop 1998;114:32-9.
Fallis DW, Kusy RP. Variation in flexural properties of photo-pultruded composite archwires: analyses of round and rectangular profiles. J Mater Sci Mater Med 2000;11:683-93.
Filmore GM, Tomlinson JL. Heat treatment of cobalt chromium alloys
of various tempers. Angle Orthod 1979;49:126-30.
Filmore GF, Tomlinson JL. Heat treatment of cobalt-chromium alloy
wire. Angle Orthod 1976;46:187-95.
Fragou S, Eliades T. Effect of topical fluoride application on titanium
alloys: a review of effects and clinical implications. Pediatr Dent
2010;32:99-105.
Frank CA, Nikolai, RJ. A comparative study of frictional resistances
between orthodontic bracket and arch wire. Am J Orthod
1980;78:593-609.
Garrec P, Tavernier B, Jordan L. Evolution of flexural rigidity according to the cross-sectional dimension of a superelastic nickel titanium orthodontic wire. Eur J Orthod 2005;27:402-407.
Gioka C, Eliades T. Materials-induced variation in the torque expression of preadjusted appliances. Am J Orthod Dentofacial Orthop
2004;125:323-8.
Gioka C, Eliades T. Superelasticity of nickel-titanium orthodontic
archwires: metallurgical structure and clinical importance. Hel
Orthod Rev 2002;5:111-27.
Goldberg AI, Burstone CJ. An evaluation of beta titanium alloys for
use in orthodontic appliances. J Dent Res 1979;58:593-600.
Goldberg AJ, Burstone CJ. The use of continuous fiber reinforcement
in dentistry. Dent Mater. 1992;8:197-202.
Goldberg AJ, Liebler SA, Burstone CJ. Viscoelastic properties of an
aesthetic translucent orthodontic wire. Eur J Orthod 2010, Dec
15. [Epub ahead of print]
Goldberg AJ, Morton J, Burstone CJ. The flexure modulus of elasticity of orthodontic wires. J Dent Res 1983;62:856-8.
Gross TS, Srinivasan S, Liu CC, Clemens TL. Noninvasive loading of
the murine tibia: an in vivo model for the study of mechanotransduction. J Bone Miner Res 2002;17:493-501.
Hamdan A, Rock P. The effect of different combinations of tip and
torque on archwire/bracket friction. Eur J Orthod 2008;30:50814.
Hensten-Pettersen A, Jacobsen N. Disintegration of orthodontic
appliances. In: Brantley WA, Eliades T, eds. Orthodontic materials: scientific and clinical aspects. Stuttgart: Thieme, 2001: 1259.
Holberg C, Holberg N, Rudzki-Janson I. Sutural strain in orthopedic
headgear therapy: A finite element analysis. Am J Orthod Dentofacial Orthop 2008;134:53-9.
House K, Sernetz F, Dymock D, Sandy JR, Ireland AJ. Corrosion of
orthodontic appliances--should we care? Am J Orthod Dentofacial Orthop 2008;133:584-92.
Houston WJB, Stephens CD, Tulley WJ, eds. A textbook of orthodontics, 2nd ed. Bristol: Wright, 1996.
Huang HH, Chiu YH, Lee TH, Wu SC, Yang HW, Su KH, Hsu CC. Ion
release from NiTi orthodontic wires in artificial saliva with vari-

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

. ,
(Kapila Sachdeva, 1989; Kusy, 1997).



(Burrow, 2009).

(binding),
,
(Thorstenson Kusy, 2002; Hamdan
Rock, 2008), (notching), ,
(Articolo ., 2000; Burrow, 2009). ,


. , - (bindingand-release)
in vitro (Swartz,
2007; Burrow, 2009). ,
SS

(Peterson ., 1982;
Kusy Whitley, 1990; Kusy Whitley, 1989; Kapila
., 1990; Cash ., 2004; Krishnan Kumar,
2004a)
.
,
,
(torque),
(
-) , , (Kapila Sachdeva, 1989;
Kusy, 1997). ,
-

(Johnson, 2003).
-
.
, (settling),

.
,

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

ous acidities. Biomaterials 2003;24:3585-92.


Huang HH. Variation in corrosion resistance of nickel-titanium wires
from different manufacturers. Angle Orthod 2005;75:661-5.
Husmann P, Bourauel C, Wessinger M, Jger A. The frictional behavior of coated guiding archwires. J Orofac Orthop 2002;63:199211.
Iijima M, Ohno H, Kawashima I, Endo K, Mizoguchi I. Mechanical
behavior at different temperatures and stresses for superelastic
nickel-titanium orthodontic wires having different transformation temperatures. Dent Mater 2002;18:88-93.
Imai T, Watari F, Yamagata S, Kobayashi M, Nagayama K, Toyoizumi
Y, Nakamura S. Mechanical properties and aesthetics of FRP
orthodontic wire fabricated by hot drawing. Biomaterials
1998;19:2195-200.
Ingram SB, Gipe DP, Smith RJ. Comparative range of orthodontic
wires. Am J Orthod 1986;90:296-307.
Johnson E. Relative stiffness of beta titanium archwires. Angle
Orthod 2003;73:259-69.
Juvvadi SR, Kailasam V, Padmanabhan S, Chitharanjan AB. Physical,
mechanical, and flexural properties of 3 orthodontic wires: an invitro study. Am J Orthod Dentofacial Orthop 2010;138:623-30.
Kapila S, Sachdeva R. Mechanical properties and clinical applications
of orthodontic wires. Am J Orthod Dentofacial Orthop
1989;96:100-9.
Kapila S, Angolkar PV, Duncanson MG, Nanda RS. Evaluation of friction between edgewise stainless steel brackets and orthodontic
wires of four alloys. Am J Orthod Dentofacial Orthop
1990;98:117-26.
Keim RG, Gottlieb EL, Nelson AH, Vogels DS 3rd. 2002 JCO study of
orthodontic diagnosis and treatment procedures. Part 1. Results
and trends. J Clin Orthod 2002;36:553-68.
Khier SE, Brantley WA, Fournelle RA. Structure and mechanical properties of as-received and heat-treated stainless steel orthodontic
wires. Am J Orthod Dentofacial Orthop 1988;93:206-12.
Kim H, Johnson JW. Corrosion of stainless steel, nickel-titanium,
coated nickel-titanium, and titanium orthodontic wires. Angle
Orthod 1999;69:39-44.
Kolokitha OE, Kaklamanos EG, Papadopoulos MA. Prevalence of
nickel hypersensitivity in orthodontic patients: a meta-analysis.
Am J Orthod Dentofacial Orthop 2008;134:e1-e12.
Krishnan V, Kumar J. Mechanical properties and surface characteristics of three archwire alloys. Angle Orthod 2004a;74:82531.
Krishnan V, Kumar KJ. Weld characteristics of orthodontic archwire
materials. Angle Orthod 2004b;74:533-8.
Kusy RP. A review of contemporary archwires: their properties and
characteristics. Angle Orthod 1997;67:197-208.
Kusy RP. Comparison of nickel-titanium and beta titanium wire sizes
to conventional orthodontic arch wire materials. Am J Orthod
1981;79:625-9.
Kusy RP, Dilley GJ. Elastic modulus of triple-stranded stainless steel
arch wire via three- and four-point bending. J Dent Res
1984;63:1232-40.
Kusy RP, Greenberg AR. Effects of composition and cross section on
the elastic properties of orthodontic arch wires. Angle Orthod
1981;51:325-41.
Kusy RP, Mims L, Whitley JQ. Mechanical characteristics of various
tempers of as-received cobalt-chromium archwires. Am J Orthod
Dentofacial Orthop 2001;119:274-91.
Kusy RP, Stevens LE. Triple-stranded stainless steel wires. Evaluation

63

/ Properties and applications of orthodontic wires

,
(0,0215
)
(Zachrisson, 2007).
(residual stresses).
-

,
.
(coated)
(fiber-reinforced). ,
,
.
, Teflon
(Elayyan ., 2008;
Elayyan ., 2010). '' ''

in vivo (Kusy, 1997).

(Lim
., 1994). 25%
33
(Elayyan ., 2008). H
- (Husmann ., 2002),

(Elayyan ., 2008).
(Elayyan
., 2010). , (

),

(Elayyan ., 2010).
,

(Zufall Kusy, 2000b; Goldberg
Burstone, 1992).
,

64

HELLENIC ORTHODONTIC REVIEW

of mechanical properties and comparison with titanium alloy


alternatives. Angle Orthod 1987;57:18-32.
Kusy RP, Whitley JQ. Thermal and mechanical characteristics of stainless steel, titanium-molybdenum, and nickel-titanium archwires.
Am J Orthod Dentofacial Orthop 2007;131:229-37.
Kusy RP, Whitley JQ. Effects of surface roughness on the coefficients
of friction in model orthodontic systems. J Biomech
1990;23:913-25.
Kusy RP, Whitley JQ. Effects of sliding velocity on the coefficients of
friction in a model orthodontic system. Dent Mater 1989;5:23540.
Kusy RP, Whitley JQ, de Ara_jo Gurgel J. Comparisons of surface
roughness and sliding resistances of 6 titanium-based or TMAtype archwires. Am J Orthod Dentofacial Orthop 2004;126:589603.
Kwon YH, Seol HJ, Kim HI, Hwang KJ, Lee SG, Kim KH. Effect of acidic
fluoride solution on beta titanium alloy wire. J Biomed Mater
Res B Appl Biomater 2005;73:28590.
Leach HA, Ireland AJ, Whaites EJ. Radiographic diagnosis of root
resorption in relation to orthodontics. Br Dent J 2001;190:16-22.
Lendlein A, Jiang H, Junger O, Langer R. Light-induced shape-memory polymers. Nature 2005;434:879-82.
Lim KF, Lew KK, Toh SL. Bending stiffness of two aesthetic orthodontic archwires: An in vitro comparative study. Clin Mater
1994;16:63-71.
Lopez I, Goldberg J, Burstone CJ. Bending characteristics of nitinol
wire. Am J Orthod 1979;75:569-75.
Meling TR, Odegaard J. Short-term temperature changes influence
the force exerted by superelastic nickel-titanium archwires activated in orthodontic bending. Am J Orthod Dentofacial Orthop
1998a;114:503-9.
Meling TR, Odegaard J. The effect of temperature on the elastic
responses to longitudinal torsion of rectangular nickel titanium
archwires. Angle Orthod 1998b;68:357-68.
Meling TR, Odegaard J. The effect of short-term temperature
changes on superelastic nickel-titanium archwires activated in
orthodontic bending. Am J Orthod Dentofacial Orthop
2001;119:26373.
Mikulewicz M, Chojnacka K. Trace metal release from orthodontic
appliances by in vivo studies: a systematic literature review. Biol
Trace Elem Res 2010;137:127-38.
Miura F, Mogi M, Ohura Y, Hamanaka H. The superelastic property of
the Japanese NiTi alloy wire for use in orthodontics. Am J Orthod
1986;90:1-10.
Morina E, Eliades T, Pandis N, Jger A, Bourauel C. Torque expression
of self-ligating brackets compared with conventional metallic,
ceramic, and plastic brackets. Eur J Orthod 2008;30:233-8.
Nakano H, Satoh K, Norris R, Jin T, Kamegai T, Ishikawa F, Katsura H.
Mechanical properties of several nickel-titanium alloy wires in
three-point bending tests. Am J Orthod Dentofacial Orthop
1999;115:390-5.
Nelson KR, Burstone CJ, Goldberg AJ. Optimal welding of beta-titanium orthodontic wires. Am J Orthod Dentofacial Orthop
1987;92:213-9.
Ogawa T, Yokoyama K, Asaoka K, Sakai J. Hydrogen absorption
behavior of beta titanium alloy in acid fluoride solutions. Biomaterials 2004;25:241925.
Oh K, Kim Y, Park Y, Kim K. Properties of super stainless steels for
orthodontic applications. J Biomed Mater Res Part B: Appl Bio-

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

EHNIKH OPOONTIKH EIEPHH

/ Properties and applications of orthodontic wires

/

(Imai ., 1998; Zufall Kusy, 2000b).
(Fallis Kusy., 2000).

(Cacciafesta .,
2008).

,
(Kusy, 1997). ,
, (Lendlein .,
2005). (Eliades, 2007).
()
- . ,

(relaxation). ,

(Goldberg ., 2010).

,
.
, .
,
,
.



.
. ,

,

. -

EHNIKH OPOONTIKH EIEPHH 2011 TOMO 14 TEYXO 1

mater 2004;69B:183-94.
Pandis N, Polychronopoulou A, Eliades T. Alleviation of mandibular
anterior crowding with copper-nickel-titanium vs nickel-titanium
wires: a double-blind randomized control trial. Am J Orthod
Dentofacial Orthop 2009;136:152.e1-7.
Partowi S, Keilig L, Reimann S, Jger A, Bourauel C. Experimental
analysis of torque characteristics of orthodontic wires. J Orofac
Orthop 2010;71:362-72.
Pelsue BM, Zinelis S, Bradley TG, Berzins DW, Eliades T, Eliades G.
Structure, composition, and mechanical properties of Australian
orthodontic wires. Angle Orthod 2009;79:97-101.
Proffit WR, Fields HM, Sarver DM. Contemporary Orthodontics. 4rth
ed. St. Louis: CV Mosby, 2007: 364, 366, 553-55.
Quint_o CC, Cal-Neto JP, Menezes LM, Elias CN. Force-deflection
properties of initial orthodontic archwires. World J Orthod
2009;10:29-32.
Redlich M, Katz A, Rapoport L, Wagner HD, Feldman Y, Tenne R.
Improved orthodontic stainless steel wires coated with inorganic fullerene-like nanoparticles of WS(2) impregnated in electroless nickel-phosphorous film. Dent Mater 2008;24:1640-6.
Reitan K. Clinical and histologic observations on tooth movement
during and after orthodontic tooth movement. Am J Orthod
1967;53:721-4.
Riley M, Bearn DR. A systematic review of clinical trials of aligning
archwires. J Orthod 2009;36:42-51.
Rucker BK, Kusy RP. Elastic flexural properties of multistranded stainless steel versus conventional nickel titanium archwires. Angle
Orthod 2002a;72:302-9.
Rucker BK, Kusy RP. Elastic properties of alternative versus singlestranded leveling archwires. Am J Orthod Dentofacial Orthop
2002b;122:528-41.
Rucker BK, Kusy RP. Resistance to sliding of stainless steel multistranded archwires and comparison with single-stranded leveling wires. Am J Orthod Dentofacial Orthop 2002c;122:73-83.
Rucker BK, Kusy RP. Theoretical investigation of elastic flexural properties for multistranded orthodontic archwires. J Biomed Mater
Res 2002d;62:338-49.
Rudolph DJ, Willes PMG, Sameshima GT. A finite element model of
apical force distribution from orthodontic tooth movement.
Angle Orthod 2001;71:127-31.
Sakima MT, Dalstra M, Melsen B. How does temperature influence
the properties of rectangular nickel-titanium wires? Eur J Orthod
2006;28:282-91.
Sebanc J, Brantley WA, Pincsak JJ, Conover JP. Variability of effective
root torque as a function of edge bevel on orthodontic arch
wires. Am J Orthod 1984;86:43-51.
Segal N, Hell J, Berzins DW. Influence of stress and phase on corrosion of a superelastic nickel-titanium orthodontic wire. Am J
Orthod Dentofacial Orthop 2009;135:764-70.
Shetty V, Caridad JM, Caputo AA, Chaconas SJ. Biomechanical rationale for surgical-orthodontic expansion of the adult maxilla. J
Oral Maxillofac Surg 1994;52:742-9.
Swartz LS. Fact or fiction: the clinical relevance of in vitro steady
state friction studies. J Clin Orthod 2007;8:427-32.
Taneja P, Duncanson MG Jr, Khajotia SS, Nanda RS. Deactivation
force-deflection behavior of multistranded stainless steel wires.
Am J Orthod Dentofacial Orthop 2003;124:61-8.
Thorstenson GA, Kusy RP. Effect of archwire size and material on the
resistance to sliding of self-ligating brackets with second-order

65

HELLENIC ORTHODONTIC REVIEW

/ Properties and applications of orthodontic wires


,
.
.


.

:
N




54124
E-mail: ntopouz@dent.auth.gr

angulation in the dry state. Am J Orthod Dentofacial Orthop


2002;122:295-305.
Van Leeuwen EJ, Kuijpers-Jagtman AM, Von den Hoff JW, Wagener
FA, Maltha JC. Rate of orthodontic tooth movement after changing the force magnitude: an experimental study in beagle dogs.
Orthod Craniofac Res 2010;13:238-45.
Verstrynge A, Van Humbeeck J, Willemsc J. In-vitro evaluation of the
material characteristics of stainless steel and beta-titanium
orthodontic wires. Am J Orthod Dentofacial Orthop
2006;130:460-70.
Walker MP, White RJ, Kula KS. Effect of fluoride prophylactic agents
on the mechanical properties of nickel-titanium-based orthodontic wires. Am J Orthod Dentofacial Orthop 2005;127:6629.
Walker MP, Ries D, Kula K, Ellis M, Fricke B. Mechanical properties
and surface characterization of beta titanium and stainless steel
orthodontic wire following topical fluoride treatment. Angle
Orthod 2007;77:342-8.
Wang Y, Jian F, Lai W, Zhao Z, Yang Z, Liao Z, Shi Z, Wu T, Millett DT,
McIntyre GT, Hickman J. Initial arch wires for alignment of
crooked teeth with fixed orthodontic braces. Cochrane Database
Syst Rev 2010;4, CD007859.
West AE, Jones ML, Newcombe RG. Multiflex versus superelastic: a
randomized clinical trial of the tooth alignment ability of initial
arch wires. Am J Orthod Dentofacial Orthop 1995;108: 46471.
Wilkinson PD, Dysart PS, Hood JA, Herbison GP. Load-deflection characteristics of superelastic nickel-titanium orthodontic wires. Am
J Orthod Dentofacial Orthop 2002;121:483-95.
Zachrisson BU. Long-term experience with direct-bonded retainers:
update and clinical advice. J Clin Orthod 2007;41728-37.
Zufall SW, Kusy RP. Stress relaxation and recovery behaviour of composite orthodontic archwires in bending. Eur J Orthod
2000a;22:1-12.
Zufall SW, Kusy RP. Sliding mechanics of coated composite wires and
the development of an engineering model for binding. Angle
Orthod 2000b;70:34-47.

Reprint requests to:


Nikolaos Topouzelis
Assistant Professor
Department of Orthodontics
School of Dentistry
Aristotle University of Thessaloniki
54124 Thessaloniki
E-mail: ntopouz@dent.auth.gr

66

HELLENIC ORTHODONTIC REVIEW 2011 VOLUME 14 ISSUE 1

You might also like