You are on page 1of 34

ARCHWIRES AND ARCHWIRE TECHNOLOGY 177

EXCELLENCE IN ORTHODONTICS 2016

Chapter

11
Archwires and Biomechanics
David Birnie
178 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Introduction
A variety of archwires is now available to the orthodontist and it can be difficult to decide which archwire has the
desired properties for a particular clinical situation.

Archwire properties
point of arbitrary clinical loading
The properties of conventional archwires are shown in Figure
10
11.1 and the terms describing the wires behaviour familiar to
yield point failure point
most.
proportional limit

Resilience is indicated by the area under the curve to the left of


Stress
the proportional limit and formability by the area under the
curve between the yield and the failure points and is the
capacity of a material to absorb energy while undergoing
elastic deformation.
0
0 Range
Springback
8 The modulus of elasticity (E) is given by the slope of the graph.
Strain Stiffness is proportional to E and springiness to its reciprocal
Figure 11.1: Properties of conventional
(1/E). Range is measured from a point 0.1% along the x-axis to
archwires the yield point.

Typically, orthodontic archwire materials are required to have the following properties:

strength formable resilience


toughness weldable aesthetics
good range springiness biocompatibility
low friction springback poor biohostability

As the cross-section of a wire is altered:

• range is affected proportionally


• springiness changes as a fourth power function
• strength changes as a cubic function.

As the length of a wire supported at one end is altered:

• range is affected as a square


• springiness as a cubic function
• strength changes proportionally

Archwire materials
Archwires may be classified by:

• material
stainless steel, Elgiloy, titanium alloys, glass, polymers
ARCHWIRES AND ARCHWIRE TECHNOLOGY 179
EXCELLENCE IN ORTHODONTICS 2016

Figure 11.2: Orthodontic archwires classified by material. Stainless steel and titanium alloys are the commonly used archwire types

• coated or non-coated
ion implantation, spray coating, sleeving
• morphology
round, rectangular, single, multistrand or braided
• composite
o made of more than one morphology (e.g.: Wonder Wire which is round in the buccal segments
and rectangular in the anterior segment)
o made of one or more materials with a mechanical join or material
having different properties in different sections of the archwire (e.g.: GAC Bioforce, Forestadent
TripleForce). TSA (Tri-Sectional Archwires from Modern Arch ) are archwires made with reduced
section stainless steel (eg: 0.018” x 0.018”) in the buccal segments and NiTi, TMA or Elgiloy
(untorqued or pretorqued 20 degrees) 0.019” x 0.025” in the anterior segments.

Evans and Durning (1996a, 1996b) have produced a useful and well-accepted classification of orthodontic alloys.

Stainless steel
Stainless steel archwires are used as working archwires. Standard grade stainless steel wires are quite adequate for
use with current fixed appliances. Highly resilient or “Australian” type wires are not required.

Elgiloy
The Elgin Watch Company developed a cobalt chromium alloy (40% cobalt, 20% chromium, 5% iron and 15% nickel)
in the 1950s which Rocky Mountain Orthodontics marketed as Elgiloy. Elgiloy’s advantage is related to the ability
to increase its resilience and strength, but not its stiffness, after formation, by heat treatment. It is now of limited
use because of the extensive use of preformed archwires and the lack of need to place complex bends in archwires;
despite this, cobalt chromium archwires are sold as Remaloy (Dentaurum) and of course Rocky Mountain
Orthodontics (Elgiloy).
180 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Titanium alloys
The following titanium alloys are used in orthodontics:

• nickel-titanium alloys
o martensitic stabilised alloys
o martensitic active alloys
• martensitic graded archwires
o austenitic active alloys
• nickel titanium copper chromium
• titanium molybdenum alloys
• titanium niobium alloys

Nickel-titanium alloys
The field of nickel titanium alloys is confusing and remains so despite profuse information within the literature.
Helpful explanations have been given by Kusy (1991, 1997), Matasa (1997) and Santoro (2001a, 2001b) and the
following terms require explanation.

Definitions
active alloy
an alloy which is capable of undergoing its anticipated phase transformation; i.e.: in which an SME can be
mechanically or thermally induced

As and Af
The start and finish temperatures at which austenite is formed in a nickel titanium alloy

austenite
historically, a specific phase of iron; also a high temperature metallurgical phase of nickel titanium with a face
centred structure (hexagonal close packed) which can transform to martensite

hyperelasticicity
a property of single crystal shape memory alloys (SMA) which show ideal elasticity such as CuAlNi, CuAlMn, CuAlBe,
CuAlNb and CuAlNiMnTi. Theoretically, archwires may be formed from the archwire may be a single crystal shape
memory material comprising a single crystal of CuAlNi, CuAlMn, CuAlBe, or CuAlNiMnTi.

martensite
historically, a metastable phase of iron resulting from the diffusionless transformation of austenite following rapid
cooling; also a low temperature metallurgical phase of nickel-titanium with a body-centred (cubic or tetragonal)
structure

Ms and Mf
The start and finish temperatures at which martensite is formed in a nickel titanium alloy

passive alloy
an alloy that is incapable of undergoing its anticipated phase transformation because extensive plastic deformation
has suppressed the transformation

phase transformation
a change in the number or character of the phases that constitute the microstructure of an alloy by a change in
crystalline structure
ARCHWIRES AND ARCHWIRE TECHNOLOGY 181
EXCELLENCE IN ORTHODONTICS 2016

pseudoelasticity Property Stainless Martensite Austenite


the mechanical analogue of steel
thermoelasticity in which, at Yield Strength 2.1 GPa 1.4-1.7 GPa 0.84 GPa
constant temperature, the Elastic Modulus 200 GPa 31-35 GPa 84-98 GPa
austenitic-martensitic phase Resistivity 76 µohm-cm 82 µohm-cm
transformation occurs with Thermal Conductivity 0.085 watt/cm-°C 0.18 watt/cm-°C
increasing applied force; the effect Magnetic Susceptibility 2.4 x 10-6 emu/g 3.7 x 10-6 emu/g
by which a material recovers the Thermal Expansion 6.6 x 10-6/°C 11 x 10-6/°C
induced plastic strain on loading.
Table 11.1: Physical properties of nickel titanium alloys as a function of their
Clinically, it is useful to think of this phase
as a local stress-related superelastic
effect generated by tying in an irregular tooth.

R structure
a metastable rhombohedral structured intermediate phase between austenite and martensite (Khier et al 1991).
The R structure always seems to be present in an austenite to martensite transformation but not always in the
reverse transformation (Bradley et al 1996). As of 2011, R phase is considered a martensitic phase which may appear
in certain testing of alloys upon incremental heating or stressing. The diffusionless reorientation of atoms in the
material must proceed on the minimal energy pathway; this may or may not include intermediate R phase existence
between austenite and martensite. Traditional evaluation of these alloys does not consider the R phase due to lack
of cohesive agreement among the academic and industrial communities. Thus only austenite and martensite are
considered, R phase being swept up in the martensite regime (Alauddin S, 2011, Personal communication).

resistivity
amount of resistance to the passage of electrical currents; because austenite and martensite have different
resitivities (see Table 11.1), the phase transformation temperatures can be inferred from the resistivity of a particular
alloy. Other methods of inferring phase transformation are radiographic diffraction and differential scanning
calorimetry

shape memory effect (SME)


the ability of a wire to return to undergo deformation and then return to its prior shape when triggered by a
mechanical or thermal stimulus. This phenomenon is a combination of thermoelasticity and pseudoelasticity.
Strictly speaking this is a one-way shape memory effect as only austenite retains the memorised shape

stabilised alloy
an alloy in which a predetermined amount of wire deformation has occurred during processing suppressing the
SME (q.v.)

stress induced martensitic transformation (SIM)


the mechanical analogue of SME. The ability of a wire to revert to its original shape when subjected to mechanical
deformation

superelasticity
the martensite-austenite phase transformations, generated by mechanical stress or heat, which result in a segment
of the loading curve for nickel titanium alloys in which stress is independent of applied strain. It only exists when
both phases of the alloy are present.

thermoelasticity
the thermal analogue of pseudoelasticity in which the martensitic phase transformation occurs from austenite as
the temperature decreases. The phase transformation can be reversed by increasing the temperature to its original
182 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

value (pseudo-shearing). In most metals and alloys, an


Mf equivalent amount of stress usually generates an
irreversible shearing deformation
Ms

transition temperature range (TTR)


Resistivity

As
the temperature range over which the alloy structure
changes from the martensitic to the austenitic phase
Af

The mode of action of nickel titanium alloys


Nickel titanium can exist in two different crystalline forms;
Temperature face-centred (hexagonal close-packed) austenite and
body-centred (cubic or tetragonal) martensite. The
Figure 11.3: Resistivity against temperature for superelastic physical difference between austenite and martensite are
nickel titanium archwires (after Santoro 2001a)
given in Table 11.1. The process of phase transformation
has been examined by measuring resistivity, using radiographic diffraction and differential scanning calorimetry
(Bradley et al 1996). A newer method, temperature modulated differential scanning calorimetry (TMDSC) gives
further information about phase transformation of nickel titanium alloys and its complexity (Brantley et al 2003).

Transformation can be induced either by change in temperature or by change in applied stress. Temperature
change induced transformation between the phases can be brought about by cooling (austenite to martensite) or
heating (martensite to austenite). From this it is apparent that martensite is a low temperature form of the alloy and
austenite the high temperature form of the alloy. Another transformation to an intermediate phase can occur
between the austenitic and martensitic phases known as the R phase; this has rhombohedral symmetry and a simple
hexagonal lattice. Thus the transformation may be austenite - R-phase - martensite and the reverse on heating. Not
all transformations pass through the R-phase however and the R-phase has a lower elastic modulus than austenite.

A review of superelastic nickel titanium archwires has been published by Waters (1992), Evans and Durning (1996)
and Santoro et al (2001a, 2001b).

Temperature transition ranges (TTRs)


In Figure 11.3, the phase changes of nickel titanium archwires are demonstrated. At low temperatures, the archwire
is completely martensite; as the temperature rises, progressive transformation into austenite occurs until finally, the
alloy is wholly composed of austenite. The period during which the phase transformation occurs (martensite to
austenite) generates the shape memory effect. For this to be clinically detectable, the Af has to be just below mouth
temperature so that the alloy is mostly in its austenitic form; however, if it is all austenite then the alloy simply obeys
Hooke’s Law (stress and strain are proportional). So ideally, the Af should be set above mouth temperature and
nickel titanium alloys should have temperature transition ranges that correspond to the temperatures of the oral
environment.

If the Af is below mouth temperature, then intraorally, the alloy is principally austenite. In order to effect
superelasticity, more martensite is necessary; this is produced by archwire deflection. This is stress induced
martensitic (SIM) transformation. Interestingly, SIM shifts the TTR to higher values since as a result of the changes
in the lattice structure, more energy is required to reverse the martensitic transformation. Accordingly, in clinical
practice one of the important characteristics of nickel titanium alloys is their stress-related Af; this is the temperature
at which thermally active wires are ‘activated’. Thus the stress-induced (pseudoelasticity) and the thermal
(thermoelasticity) properties of nickel titanium alloys are related (Santoro and Beshers 2000).

In laboratory studies, the type of loading should simulate the clinical situation as closely as possible. At the moment
the three bracket bending test is the most appropriate (Santoro et al 2001b).
ARCHWIRES AND ARCHWIRE TECHNOLOGY 183
EXCELLENCE IN ORTHODONTICS 2016

0.016" wire: load at 1.5 0.016" wire: difference in 0.016 x 0.022" wire: load 0.016 x 0.022" wire:
mm activation load between 0.5 mm and at 1.5 mm activation difference in load
1.5 mm activation between 0.5 mm and 1.5
mm activation
A Company Align 153 100 340 200
Ormco CuNiTi 35 17 17
Ormco CuNiTi 40 3 3

Table 11.3: Load characteristics of different nickel titanium wires (in grams)

The force generated by superelastic behaviour varies with temperature. In a paper by Filleul (1997), three out of
four tested “superelastic” wires showed superelasticity at 22°C, two out of four at 39°C and none of the four at 44°
C. The start of the superelastic plateau occurs at different force levels according to temperature; lower temperatures
generate a plateau at a lower force. This is an illuminating paper which repays careful reading. This variability has
been re-emphasised by Nakano et al (1999) who tested 42 brands of nickel titanium wires and found large
differences in the loads exerted in three-point bending tests as shown in Table 11.3 as well as in the plateau
behaviour of different archwires. Similar variability has been found by Gurgel et al (2001). Bolender et al (2010)
have suggested that most NiTi wires do not exhibit in torsion the superelastic effect traditionally described in
bending and that the optimal constant moments necessary to gain third-order control of tooth movement early in
treatment are not present in a preadjusted edgewise-rectangular NiTi archwire system. A braided stainless steel
rectangular archwire displayed better torsional behaviour at 35°C than most NiTi archwires of the same dimensions.
It is important to remember that:

• thermoelastic wires make archwires easier to tie in


• superelasticity in the mouth comes from stress induced martensitic transformation

Differential scanning calorimetry (DSC) method (measures the amount of energy (heat) absorbed or released by a
sample as it is heated or cooled) was used by Ren et al (2008) to study the phase transformation temperatures and
the phase transition processes of nine commonly used clinical NiTi alloys in China (0.016” and 0.016” x 0.022”) (Figure
11.2). The Af of the Smart, Ormco and 3M NiTi round wires were lower than room temperature and no phase
transformation was detected during oral temperature indicating that these archwires do not possess shape
memory. No phase transformation was seen for Youyan 1 round wire from -80° to +80°. For the rectangular wires,
the Af of Smart, L&HR and Youyan 2 were close to mouth temperature whereas at room temperature they were
mixture of martensite and austenite suggesting these wires had a shape memory effect. No phase transformation
for the rectangular Youyan 1 was seen from -80° to +80° suggesting this wire had no shape memory effect. The
0.016” x 0.022” Damon CuNiTi wire existed as a mixture of martensite and austenite at room temperature but its Af
was much higher than mouth temperature at 61.5°C and was therefore almost completely martensitic at mouth
temperature. One slightly puzzling feature about this paper is that we were not aware that a 0.016” x 0.022” Damon

Product Dimension(s) Manufacturer Abbreviation


Smart SE 0.016” Beijing Smart Tech Co Ltd Smart
Kleen Pak Ni-Ti 0.016” Ormco Ormco
CY16 0.016” Grikin Advanced Materials Co Youyan 1
Nitinol heat activated 0.016” 3M Unitek 3M
Smart SE 0.016” x 0.022” Beijing Smart Tech Co Ltd Smart
Damon CuNiTi 0.016” x 0.022” Ormco Damon
L&H Titan 0.016” x 0.022” Tomy International L&HR
CF1616 0.016” x 0.022” Grikin Advanced Materials Co Youyan 1
RF1616 0.016” x 0.022” Grikin Advanced Materials Co Youyan 2

Table 11.2: Types of NiTi wire used in the study by Ren et al (2008)
184 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

CuNiTi wire has ever been made. The somewhat strange results reported by this paper are as a result of differing
test data definitions related to the existence of the R-phase.

Three types of nickel titanium alloys are available now:

Martensitic stabilised alloys


An example of these alloys is Nitinol. Nitinol undergoes 8-10% deformation during processing which suppresses
the shape memory effect. This alloy has a mainly martensitic structure at room temperature. The alloy delivers one-
fifth to one-sixth the force per unit activation thus giving a light, continuous and linear force.

Austenitic active alloys


The introduction of super-elastic nickel titanium alloys (Miura et al 1986) has been an important step forward in wire
technology. Earlier nickel-titanium alloys were work hardened and possessed a high elastic limit and a low elastic
modulus giving excellent spring back. Super-elastic nickel titanium wires have excellent springback, shape memory
and superelasticity. These archwires have primarily austenitic grain structure with some martensitic Ni-Ti present.
A phase called the R structure, which is an intermediate metastable rhombohedral phase, may exist in the
transformation between austenite to martensite.

These alloys are, by definition, predominantly in the austenitic phase at room temperature and are engineered not
to undergo thermal phase transformation at mouth temperature. If the austenitic transformation temperature is
below mouth temperature, then a stress induced martensitic transformation can however be induced by
deformation. This phenomenon is variously called pseudoplasticity or pseudoelasticity (a subset of super elasticity)
because of the flat force deactivation curve. Since austenite has a higher elastic modulus, on loading, the slope of
the graph starts with a slope three times that of martensitic stabilised alloy; this then gives way to plateau behaviour
and finally a further positive slope characteristic of martensitic nickel titanium.

Both types of active alloys can exhibit this type of stress-induced transformation and thus be superelastic.

Austenitic active alloys, then, deform elastically within the austenite phase when an initial strain is applied to them.
Strain induces martensitic transformation during which increasing strain is not accompanied by an increase in stress.
Once martensitic transformation is complete, then the material again behaves elastically within the martensitic
phase.

Martensitic active alloys


Martensitic active archwires are, by definition, predominantly martensitic at room temperature and are
thermoelastic – often referred to as thermal wires. Thermoelastic archwires change their properties according to
temperature such as martensitic active alloys. Bishara et al (1995) have produced a useful list of the ideal properties
of thermoelastic archwires. These are:

• highly ductile at room temperature


• instantaneous activation at mouth temperature
• ability to develop forces that will produce tooth movement
• once fully activated, the wire is not
Phase Start/Finish Temperature °C further activated by the heat of the mouth
Martensitic start 14 • a narrow temperature transition range
finish 7 such that the wire is highly ductile at room
Austenitic start 34 temperature and highly active at mouth
finish 43 temperature

Table 11.4: Example start and finish transformation


temperatures for martensitic active alloys
ARCHWIRES AND ARCHWIRE TECHNOLOGY 185
EXCELLENCE IN ORTHODONTICS 2016

For two wires of equal dimensions, one composed of austenite would have an initial stiffness four times greater
than one composed of martensite.

GAC's Sentalloy wire is an example of a martensitic active alloy. It is subject to a thermoelastic effect triggered by
the temperature of the mouth that raises the martensitic alloy through its critical temperature range causing the
alloy to transform to austenite and return to its original shape and stiffness. Technical improvements in the ability
to control precisely the composition of alloys have produced alloys in which the start and finish of martensitic and
austenitic transformations may be carefully controlled. Typical values for this type of alloy are given in Table 11.4.

Thus these archwires are ligated, in theory, in their malleable martensitic form and then undergo austenitic
transformation as the body heat increases. This induces a thermally activated shape memory effect which returns
the archwire to its original (austenitic) archform.

The shape memory effect has been investigated by Hurst et al (1990). This demonstrated 89% to 94% recovery for
most archwires. Sentinol (heavy) showed a mean recovery of 41.3% probably because its transition temperature
range was close to room temperature. This paper, however, investigated the laboratory behaviour of these wires
and a study that is closer to the clinical situation is necessary before firm conclusions can be drawn.

NeoSentalloy is similar to Sentalloy but has greater heat sensitivity and a higher shape memory effect. Sentalloy
Bioforce gives differential forces rising from 100g anteriorly in the midline to 320 g in the molar region as shown in
Figure 11.4. A similar alloy is produced by Forestadent (TripleForce).

Tonner and Waters (1994) have compared various superelastic nickel-titanium wires using a three-point bending
test.

In the articles, they make several important points:

• wire of the same diameter from different manufacturers has highly variable performance. This varies
at 35°C between 20 and 130 cN/mm
• superelastic wires have to be deflected at least 2 mm before exhibiting plateau behaviour
• superelastic wires show significant temperature sensitivity over the range form 5°C to 30°C

Nickel Titanium Copper Chromium Alloys (CuNiTi)


The addition of copper to nickel titanium alloys increases strength, reduces hysteresis (the energy lost between
activation and deactivation) and allows greater precision in the setting of the austenitic transformation
temperature. The addition of copper, however, increases the transformation temperature to above that of the oral
cavity and to compensate for this, 0.2 to 0.5% chromium is added to reduce the transformation temperature. CuNiTi
was originally produced with four different austenitic
350
transformation temperatures covering both
pseudoelastic and thermoelastic archwires. 300

250
CuNiTi type 1 200
• austenitic transformation temperature 15°C
grams

150
• pseudoelastic
100
• infrequently used
50
• very heavy forces
• few clinical indications 0
0 9 18 27 36 48
• not commercially available mms from midline

Figure 11.4: GAC Sentalloy Bioforce wire


186 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

CuNiTi type 2
• austenitic transformation temperature 27°C
• pseudoelastic
• in patients
o who have an average or high pain threshold
o with normal periodontal health
o where rapid tooth movement is required and the forces generated by the archwire are constant

CuNiTi type 3
• austenitic transformation temperature 35°C
• thermoelastic
• in patients
o who have a low to normal pain threshold
o whose periodontium is normal to compromised
o where relatively low forces are desired

CuNiTi type 4
• austenitic transformation temperature 40°C
• thermoelastic
• in patients
o who are sensitive to pain
o who have compromised periodontal conditions
o where tooth movement is slowed down
o good as initial rectangular wire

Whilst the ability to later transition temperatures is important, contemporary CuNiTi wires have a variety of
transition temperatures. The addition of copper makes for a more consistent wire; Ormco have released a 0.013”
CuNiTi wire is claimed to have 65% of the force of an 0.014” CuNiTi wire.

Pandis et al (2009) investigated whether 0.016” 35° CuNiTi or 0.016” NiTi archwires resolved mandibular anterior
crowding at different rates in a double blind randomised control trial. Using In-Ovation-R active self-ligating
brackets with an 0.022” slot, they found that wire type had no effect on the rate of resolution of anterior mandibular
crowding (129.4 vs 121.4 days). Patients with an irregularity index < 5 took a shorter time to resolve anterior
mandibular crowding than those with an irregularity index > 5 (138.5 vs 113.1 days) – not surprisingly.

Clinical effectiveness of superelastic wires


There is little published work on the effectiveness of superelastic nickel-titanium wires. O'Brien et al (1990) showed
no significant difference in contact point alignment between Nitinol (MSA) and Titanol (AAA) archwires over a five-
week period. A later paper (West et al 1995) compared 0.0155" multistrand stainless steel round wire with an 0.014"
NiTi superelastic wire. The superelastic wire gave improved alignment in the lower labial segment only and with
doubtful cost benefits.

A further paper (Evans et al 1998) compares the performance of three aligning archwires: an 0.016" x 0.022" active
martensitic wire (American Orthodontics Titanium Heat Memory Wire), a graded 0.016" x 0.022" active martensitic
wire (GAC Bioforce Sentalloy) and a 0.015" multistrand wire (Dentaurum Dentaflex). Patients were included in the
study if they were having upper and lower fixed appliances and under 18 years of age. Impressions were taken at
the prealignment phase, 4 weeks into treatment and 8 weeks into treatment and the study models analysed at each
time period using a three dimensional Reflex microscope. Although visual inspection of the data would suggest
that the active martensitic wires were superior in obtaining arch alignment, this apparent finding was not supported
ARCHWIRES AND ARCHWIRE TECHNOLOGY 187
EXCELLENCE IN ORTHODONTICS 2016

by statistical analysis which found no difference Unloaded 1 mm Load 6 mm Load


between the performances of the three wires. Yet GAC Neo-Sentalloy 9-22 9-28 9-28
again, researchers have found it impossible to Ormco CuNiTi 27° 16-31 4-33 4-33
demonstrate clinically the supposed benefits of Ormco CuNiTi 35° 7-35 7-37 23-41
modern nickel titanium alloys. Some of the reasons Ormco CuNiTi 40° 13-38 21-38 21-38
for these findings are given by Riley and Bearn 3M Unitek Nitinol Heat 8-38 4-38 4-38
(2008). Activated

Table 11.5: Stress related TTRs for various archwires under two
At the present time, some authors suggest chilling different load conditions (Santoro 2000)
superelastic archwires before insertion to allow the
tying in of thicker archwires. This can be done with an ‘Ice’ aerosol or by leaving the archwires in a refrigerator or
freezer. It may be that such wires are best tied in with steel ligatures rather than elastomerics. Additionally, some
orthodontists recommend that their patients use daily warm water rinses while superelastic archwires are in place.
In reality, none of these measures seems necessary.

The effect of stress


Nickel titanium archwires are commonly used during the aligning phase of orthodontic treatment; we are therefore
particularly interested in the performance of these wires in situations of maximum load.

Segner and Ibe (1995) threw some doubt upon the clinical relevance of superelasticity when they tested sixteen
superelastic wires from nine manufacturers. Several of the archwires tested showed no pseudo-elastic properties.
None of the materials had a plateau that extended below 0.28 mm and 0.5 mm was considered a good value.

For small archwire deflections therefore, none of the materials will behave superelastically. It is known that the
presence of stress increases the Af value. With large deflections, the considerable load applied to the wire may result
in localised SIM formation thus elevating the TTR. The TTR may move out of the oral temperature range thus
rendering the archwire non-superelastic at mouth temperature. Santoro and Beshers (2000) compared five different
archwires with minimum (1 mm) and maximum (6 mm loading). The specimens were placed in a temperature bath
and the temperature raised from 4°C to 60°C and the phase determined by measuring resistivity. The results are
shown in Table 11.5. It can be seen that Ormco’s CuNiTi 27° and 3M Unitek’s Nitinol Heat Activated provide a wide
TTR, corresponding to mouth temperature changes, and unchanged by maximum loading.

The effect of temperature


Whilst thermoelastic archwires demonstrate exciting and unusual properties in relation to their physical behaviour,
the clinical relevance of this is a matter of conjecture.

Santoro et al (2001b) suggest that at least 2 mm of deflection are necessary for the formation of SIM in austenitic
wires; below 2 mm may result in a higher force delivery due to the continued presence of the austenitic phase. The
disappointing clinical results of superelastic archwires are attributed to the fact that many superelastic archwires:

• do not exhibit plateau behaviour


• require excessive deflection to achieve plateau behaviour
• deliver excessive force on the plateau
• a combination of the above factors

In several articles by Meling and Ødegaard (1998a, 1998b, 1998c), the effect of temperature changes on
thermoelastic wires has been investigated. These papers looked at the effect of temperature on the elastic
properties of thermoelastic wires (1998a), the effect of temperature on longitudinal torsion (1998b) and the effect
of temperature on bending (1998c).
188 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

In 1998a, Meling and Ødegaard confirmed Segner and Ibe’s suspicions when they found that only four out of eight
superelastic nickel-titanium wires showed superelastic behaviour when deactivated from 25 degrees of torsional
twist. Only one wire had a wide plateau at twist angles present in conventional prescriptions. The other wires
demonstrated deactivation plateaus when deactivated from 45 to 60 degrees of torsion which is outside the range
employed in orthodontics; in addition, the thermal responsiveness of the wires was variable between wires.

In 1998b, the effect of short-term temperature changes was investigated in archwires in torsion. Meling and
Ødegaard found that the torsional stiffness of superelastic nickel titanium wires was significantly influenced by the
short-term application of hot (80°C) or cold (10°C) water. The changes induced by hot water were transient but cold
water induced a reduction in torsional stiffness which persisted for at least five minutes after the application of the
water. Repeated application of cold water resulted in incremental reductions in torsional stiffness which was still
present 60 minutes later.

Similar effects were found in bending (1998c) although martensitic stabilised archwires, which are relatively
temperature insensitive, showed minimal changes.

Rucker and Kusy (2002a) compared the elastic flexural properties of multistranded stainless steel wires with those
of conventional (stabilized martensitic) nickel titanium archwires. The nickel titanium archwires showed non-linear
elasticity thus underestimating the elastic properties of strength and range. Multistranded stainless steel wires did
not match the strength and range of conventional NiTi alloys; however, the paper suggests that it should be possible
to enhance the performance of multistranded stainless steel wires by decreasing their yield strengths.

Rucker and Kusy (2002b) also compared resistance to sliding for multistranded stainless steel and single stranded
wires. They found that there was almost no difference between the kinetic coefficients of friction for stainless steel
wires, regardless of configuration, in the passive region (<θc). In general, nickel titanium archwires had higher
coefficients of kinetic friction. In the active region (>θc), the multistrand wires behaved as follows compared with
single strand wires: coaxial wires had low friction, three-strand round and eight strand rectangular medium friction
and three strand rectangular wires had high friction.

Rucker and Kusy (2002c) have shed some light on why nickel titanium archwires are instinctively preferred by
orthodontists in the early stages of treatment and yet it has been difficult to demonstrate the benefits of nickel
titanium archwires over other initial aligning archwires scientifically. Rucker and Kusy argue that an initial aligning
archwire requires:

• high strength to prevent permanent deformation


• low stiffness to deliver low forces on activation
• high range to maximize activations

The authors tested three groups of archwires: wires with similar overall wire diameters (0.018"); group wires with
stainless steel cores of polymer coated wires with single-stranded SS wires; and braided wire configurations with
corresponding single stranded wires. These groups were compared with 0.016" nickel titanium archwire and the
elastic property ratios calculated. The behavior of superelastic nickel titanium wires can be divided into three
distinct regions: the A (austenitic) region in which the elastic properties can be predicted from linear elastic models,
where the E measurements for these archwires typically vary from 33 to 55 GPa; the SP (superelastic plateau) region
in which wires deliver a nearly constant force over a range of activation, the extent of which depends on the initial
content of the A phase; and the M (martensitic) region in which the magnitudes of E should be less than those of
the A region and typically equal 31 to 35 GPa. Here, the elastic behavior can also be predicted by a linear elastic
model. In this study, the elastic properties of single stranded nickel titanium wires, were superior to all the other
levelling wires. This outcome is independent of whether the wires are clinically operating in the austenitic region
or in the superelastic plateau region.
ARCHWIRES AND ARCHWIRE TECHNOLOGY 189
EXCELLENCE IN ORTHODONTICS 2016

Interestingly, the paper also points out that bevelling and rounding of rectangular archwires can impact the flexural
properties. For example, an 8% decrease in cross-sectional area (i.e., 2% from each corner) results in nearly a 20%
decrease in stiffness.

In contrast, Badran et al (2003) used photoelastic models to simulate the shear stresses involved in aligning
moderately crowded lower incisors and a palatally displaced upper canine using 0.015" and 0.0175" multi-strand,
0.014" and 0.016" martensitic stabilised (classic) nickel titanium and 0.014" and 0.016" austenitic active (stress
induced superelastic) nickel titanium archwires. The martensitic stabilised nickel titanium archwires delivered
forces 15-20% higher than the forces produced by the multi-strand and martensitic stabilised archwires which
delivered similar forces. This paper then again shows that it is difficult to conclusively demonstrate the superiority
of nickel titanium archwires over multi-strand archwires. In contrast, Weilland (2003) in a rather ingenious clinical
experiment using 0.016" stainless steel solid archwires and 0.016" GAC (blue) Sentalloy superelastic alloy. Weilland
found that the superelastic archwire moved teeth significantly further than stainless steel (median 3.5 mm
compared with 2.3 mm on an activation of 4.5 mm) over a period of 12 weeks. However, this came at the price of
statistically significantly more root resorption in the teeth treated with the superelastic wire.

Gatta et al (2011) investigated load-deflection characteristics of superelastic and thermal nickel titanium archwires
using a three-point bending technique. The authors concluded that thermal archwires demonstrated significantly
lower working forces than superelastic archwires at both 2 mm and 4 mm deflections. Increase in wire size produced
an increase in working force. In the unloading phase, a statistically significant decrease in force with increasing wire
deflection occurred for all archwire types. Superelastic wires at 2 mm of deflection showed curves with a smaller
hysteresis than thermal wires of the same size. At 4 mm deflection, the superelastic and thermal wires of the same
size demonstrated similar behaviour, characterized by wide hysteresis curves but with different force levels.

Copper aluminium nickel alloys (CuAlNi)


These are hyperelastic wires made from single crystal shape memory alloys (SMA). In contrast, nickel titanium
archwires are polycrystalline. Single crystal SMAs are of the from CuAlX where X is Ni, Fe, Co or Mn although in
orthodontics the CuAlNi alloy is used with about 14.5% aluminium and 4.5% nickel. CuAlNi has much greater strain
recovery than NiTi alloys and does not exhibit gradual change due to repeated cycling as there are no grain
boundaries. The advantages of hyperelastic wires over superelastic wires are:

• greater than 9% strain recovery which is three times greater than nickel titanium. This means that
the plateau section of the stress strain graph is much longer for hyperelastic alloys
• true constant force deflection. Unlike polycrystalline materials which reach their strain/stress plateau
strength in a gradual fashion and maintain an upward slope when deformed further, hyperelastic
SMA materials have a very sharp and clear plateau strain/stress that provides a truly flat spring rate
when deformed up to 9%. The stress level at which the plateau occurs depends on the temperature
difference between the transformation temperature and the loading temperature. Copper-based
hyperelastic single crystal SMAs exhibit generally lower stress levels than titanium-based alloys. In
fact, because the stress-induced martensite transformation is complete, the stress plateau can be
near zero or as large as several hundred megapascals depending on composition and temperature.
This adjustable nature of hyperelastic SMAs allows greater versatility in clinical applications. In
addition, some single crystal SMAs (eg: CuAlBe) exhibiting hyperelasticity benefit from a second
stress plateau which can increase the total recoverable strain to 22%
• very narrow loading-unloading hysteresis. As a result, there is substantially the same constant force
spring rate during both loading and unloading. This means that the force that the clinician feels as
the archwire is ligated closely approximates to the force applied to the tooth as the archwire unloads.
The stress level at which the plateau occurs depends on the temperature difference between the
transformation temperature and the loading temperature.
190 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

20 • recovery which is 100 percent


18 repeatable and complete. One of the
16 disadvantages of polycrystalline SMA
14 materials is that “settling” that occurs as
12 the material is cycled back and forth; the
10 settling problem has required that the
N

8
material be either “trained” as part of the
6
manufacturing process
4
2
• an intrinsic hyperelastic property. NiTi
0
SMA can be conditioned, through a
0.014 0.016 0.014 x 0.025 0.016 x0.025 0.018 x 0.025 combination of alloying, heat treatment
CuNiTi CuAlNi
and cold working, to have superelastic
properties. A single crystal CuAlNi SMA has
intrinsic hyperelastic properties: a crystal of
Figure 11.5: Binding test CuNiTi versus CuAlNi
CuAlNi is hyperelastic immediately after
being formed (pulled and quenched) with no further processing required
• other properties unlikely to be of benefit in orthodontics are very low yield strength when martensitic
and ultra-low transition temperature (can be close to absolute zero)

Manufacturing hyperelastic alloys


Single crystals cannot be processed by conventional hot or cold mechanical formation without breaking the single
crystallinity; a special procedure is therefore required to shaping single crystals in the process of growth as the
crystal is pulled from the melt, resulting in the finished shape.

Single crystal SMA is made in a special crystal-pulling apparatus. A seed of the desired alloy is lowered into a crucible
containing a melted ingot of the alloy composition, and gradually drawn up. The top surface of the melt contains a
die (of the desired cross-sectional shape) that forms the shape of the crystal as it grows. Surface tension pulls the
melted metal along with the seed. The rising column cools as it leaves the surface of the melt. The rate of drawing
is controlled to correspond with the rate of cooling so that a solid crystal is formed at a region that becomes a
crystallization front. This front remains stationary while the crystal, liquid below and solid above, travels through it.
This procedure generally is known as the Stepanov method of making single crystals.

The shape of the archwire is set by heating a single crystal shape memory alloy material to an annealing temperature
(Ta); and driving the single crystal shape memory alloy material at the annealing temperature and a shaping form
together and into a quenching medium; wherein the heating and driving steps are performed in less than about a
minute.

Possible clinical advantages of hyperelastic archwires


The hyperelastic SMA archwires may simplify treatment of malocclusions by combining the functions of different
archwires currently required for use during different phases of treatment. The material properties of Cu based single
crystal wires could merge much of the initial and intermediate phases of treatment, thus reducing treatment time.

In comparison to NiTi and TMA, single crystal SMA hyperelastic archwires may provide improved patient comfort
(low force), osteocompatibility for efficient bone remodelling (a constant force which can be precisely set),
resistance to permanent set for severe malocclusions (greater superelasticity), clinical ease of use (no hysteresis),
resistance to cyclic failure (longer fatigue life) and case efficiency (lower frictionless binding). Because a hyperelastic
SMA wire behaves as a smaller wire than an equivalently-sized NiTi counterpart (eg: exerts forces equivalent to those
typically seen in smaller NiTi wires), there is an opportunity to introduce larger rectangular wires (torque) much
earlier in the treatment, effectively merging the early stages of treatment). For example, a 0.027" round CuAlNi wire
exhibits a force equivalent to that of a 0.018 NiTi wire and the greater superelastic range accommodates much more
recoverable strain, reducing likelihood of a premature set in the wire (which effectively renders it inactive until the
ARCHWIRES AND ARCHWIRE TECHNOLOGY 191
EXCELLENCE IN ORTHODONTICS 2016

next visit). The net effect of these improvements in 450

wire performance may be a reduction in the quantity 400


350
of wires required for clinical treatment compared with
300
traditional NiTi wires. If hyperelastic wires reduced a
250
traditional six wire sequence to a four or three wire
200
case, then the frequency of visits and overall treatment

grams
150
time may be reduced with benefits to both patient and
100
clinician.
50
0
Titanium molybdenum alloys (TMA) Damon CuNiTi Heat activated NiTi Supercable
Stabilised beta-titanium alloys (TMA) contain 80% NiTi
0.012" 0.013" 0.014" 0.016" 0.018" 0.020"
titanium as well as 10% molybdenum, 6% zirconium
and 4% tin. TMA is twice as stiff as martensitic Figure 11.6: Unloading forces for various types and dimensions
stabilised nickel titanium alloys and a third as stiff as of archwires with 3 mm of activation. Supercable delivers approximately
stainless steel. It is formable and weldable. Proffit one third of the force of an equal diameter round nickel titanium archwire
(after Berger 2008)
(1986) cites the use of TMA wires to provide root
paralleling at the end of treatment; these archwires are probably more useful for the delivery of localised tip and
torque. It is therefore suited to obtaining more precise individual tooth movements such as in-out, tip, torque and
rotation, particularly at the end of treatment.

Originally, titanium molybdenum alloys were protected by patent; however, this expired a few years ago and several
variants of the alloy are now available. These have been described by Verstrynge et al (2006) as follows:

• Ti-11.5, Mo-6, Zr-4.5,Sn


o beta III – β-alloy
• Ti-3, Al-8, V-6, Cr-4, Mo-4, Zr
o beta C – β-alloy
• Ti-6, Al-4, V
o α-β alloy
• Ti-45 Nb
o titanium niobium

Twelve TMA archwires were tested and most were from the first category (Beta III). The archwires differed in their
elastic modulus, hardness, yield strength and ductility. One archwire (α-β alloy) demonstrated the highest elastic
modulus, the highest yield strength and the highest hardness. A Beta III alloy demonstrated the greatest ductility.
TMA has a higher rigidity and lower recoverable deformation than nickel titanium allows. Laheurte et al (2007) have
described a method of improving the pseudoelasticity of TMA by controlling the grain size. This produces an
archwire with mechanical properties midway between martensitic stabilised NiTi and active NiTi alloys. The
modified TMA does not however exhibit plateau behaviour.

Titanium niobium finishing arches (TiNb/FA)


Titanium niobium archwires are available and are a variant of TMA. These wires have 60% of the stiffness of TMA
and are easy to bend. They are recommended by the manufacturer for use as finishing archwires.
192 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Archwire morphology and


treatments
Supercable
Supercable 1 (Strite Industries Ltd) is a seven strand
super elastic nickel titanium coaxial wire. It is claimed
Figure 11.7: The Hills Dual-Geometry wire has a square anterior to have one fifth the force of an equal diameter round
section and polished round posterior section to allow space closure in the single strand nickel titanium wire and one third of the
posterior segments while maintaining torque control anteriorly. force of an equal diameter round stainless steel (Figure
11.6). This develops very low forces in the initial stages
of treatment, approximately one third of the force of an equal diameter round nickel titanium wire, as described by
Berger (2008).

Dual dimension archwires


Dual dimension archwires normally have a rectangular section in the anterior region and round wire posteriorly.
Wonder Wire (Wonder Wire Corporation) was the first archwire to have this feature but the company no longer
seems to be in business; we have used this wire and while an interesting diversion, we were not sufficiently enthused
by it to become a regular part of our archwire sequencing. The Hills Dual-Geometry 2 wire (see Figure 11.7) has a
square anterior section and a round posterior section – the archwire dimensions are therefore 0.018” x 0.018”
x0.018” for 0.018” slots and 0.021” x 0.021” x 0.021” for 0.022” slots and the concept of dual dimension wires remains
an interesting one.

Reverse curve nickel titanium archwires


The use of reverse curve nickel titanium archwires is an effective method of dealing with reluctant overbites.
Although slightly awkward to place and requiring some 3 months or so to act, they are a useful adjunct to treatment.
The 0.019" x 0.025" size is preferred. If left in place too long the labial segment develops a semilunar curve; this,
however, is easily removed with a flat archwire.

Pretorqued nickel titanium archwires


It is now possible to purchase nickel titanium archwires with 20° of torque in the anterior segment only. These
archwires are available as flat archwires and archwires with a reverse curve and are known as SET (SuperElastic
Torqued) archwires and are available from several orthodontic suppliers.

Mittal et al (2012) measured the torque transmission from pretorqued NiTi wires and found that this was highest
with high torque brackets (MBT). However the magnitude of the torqueing couple produced was low at between
3.3 N-mm and 3.6 N-mm on the incisors. This is lower than the reported clinically effective torque moments for
lingual root movement of an upper central incisor of 5–20 N-mm which casts doubt on the clinical effectiveness of
these archwires.

Ion implantation
Ion implantation is a technique used to modify surfaces exposed to corrosion or wear. Nitrogen ions are produced
in a source chamber from collisions between atoms and electrons emitted from a heated filament. The beam of
ions is accelerated and sent into an evacuated implantation specimen where it penetrates the specimen (Mizrahi et
al 1991).

1
SPEED System Orthodontics, Strite Industries Ltd, 298 Shepherd Avenue, Cambridge, Ontario, N3C 1V1 Canada
2
SPEED System Orthodontics, Strite Industries Ltd, 298 Shepherd Avenue, Cambridge, Ontario, N3C 1V1 Canada
ARCHWIRES AND ARCHWIRE TECHNOLOGY 193
EXCELLENCE IN ORTHODONTICS 2016

GAC (Spire Ion Implantation) and Ormco (TMA Low Friction) have produced wires treated by ion implantation to try
and reduce friction. Ormco claims a reduction in friction of 54% for Low Friction TMA.

Ryan et al (1997) investigated the effects of ion implantation on various wires in vitro and determined that stainless
steel produced the least frictional force, followed by ion implanted nickel titanium, ion implanted beta titanium,
untreated nickel titanium and untreated beta titanium. The interesting point about this article is that stainless steel
still performed better than nickel titanium alloys whether ion implanted or not.

Kula et al (1998) have cast some doubts on this and other studies in reporting that the rate of space closure in vivo
was not different for implanted and unimplanted beta titanium wires and furthermore that the rate of space closure
was not significantly different between beta-titanium and stainless steel wires. The brackets used were not
implanted. It may be that both archwire and bracket surfaces need to be implanted in order to produce significant
reductions in friction. This used a split archwire design where one side only of a 0.019" x 0.025" TMA archwire was
implanted with nitrogen ions by Spire Corporation. The computed monthly rate of space closure for TMA was 0.12
to 2.46 mm per month compared with 0.76 to 1.75 mm per month for stainless steel.

Aging and archwire fracture


Eliades and Bourauel (2005) have published an interesting review on the effect of intraoral aging on orthodontic
materials. The mechanism whereby archwires may be affected by aging are the effects on:

• friction
• force delivery
• superelasticity and
• fracture resistance

Age changes on the surface of the bracket slot or archwire due to calcified biofilm may affect the friction between
the bracket slot and archwire. In addition, the actual structure of the archwire or bracket may be affected thus
altering their mechanical properties.

Ni-Ti archwires show pitting, corrosion and biofilm deposits as a result of aging; this however is very difficult to
simulate in vitro and considerable caution should be taken in extrapolating laboratory findings to the clinical
situation.

It has been speculated that the superelasticity of NiTi wires may be affected by intraoral aging although there is no
evidence to support this; it is known however that significant temperature changes can adversely affect the
superelastic properties of NiTi wires.

The fracture resistance of NiTi wires is however significantly reduced by intraoral aging and that this affects NiTi
more than stainless steel or TiMb wires.

Zinelis et al (2007) have studied the fracture mode of nickel titanium archwires because retrieved archwires have a
much higher fracture incidence in contrast to laboratory experiments which showed that these wires were
practically unbreakable. Most fractures were located in the posterior region of the arch, probably because of high-
magnitude masticatory forces. Brittle fracture without plastic deformation was observed in most specimens
regardless of the type of nickel titanium archwire. There was no increase in the hardness of the intraorally exposed
specimens regardless of wire type which suggests that hydrogen embrittlement was not the cause of fracture.
194 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Archwire sequences
Is there evidence to recommend specific archwire sequences?
Mandall et al (2006) have undertaken a randomised clinical trial on archwire sequences which measured time taken
to get into an 0.019” x 0.025” working archwire, discomfort and root resorption. The archwire sequences were:

• 0.016” NiTi, 0.018” × 0.025” NiTi and 0.019” × 0.025” stainless steel
• 0.016” NiTi, 0.016” stainless steel, 0.020” stainless steel and 0.019” × 0.025” stainless steel
• 0.016” × 0.022” CuNiTi, 0.019” × 0.025” CuNiTi and 0.019” × 0.025” stainless steel

The authors found no difference in discomfort or amount of root resorption but the sequence with four archwires
took a greater number of visits (although not a statistically significant time) to reach 0.019” × 0.025” stainless steel
although this was not explained by the increased number of archwires. Although these are not archwire sequences
we use, this is an interesting paper that raises many questions about our conceptions of archwire sequences.

Abdelrahmen et al (2015) compared the alignment efficiency of three different nickel titanium archwire using Little’s
Irregularity Index. The archwires were:

• 3M Unitek conventional NiTi


• 3M Unitek superelastic NiTi
• 3M Unitek thermoelastic NiTi

This study clearly demonstrated demonstrates that there is no significant difference among the three types of NiTi
wires (conventional, superelastic, thermoelastic) in terms of alignment efficiency or the time required to achieve
complete alignment during the initial aligning stage.

In a Cochrane Library systematic review by Jian et al (2013), the review concluded that was no reliable evidence
from the trials included in the review that any specific initial arch wire material is better or worse than another with
regard to speed of alignment or pain. There was no evidence at all about the effect of initial arch wire materials on
the important adverse effect of root resorption.

We know that for a given nominal size, different nickel titanium archwires have different properties. Our preferred
high technology archwire is copper nickel titanium. The manufacturing process for nickel titanium archwires
struggles to keep a consistent temperature transition range and so for any nickel titanium archwire, variability of
physical properties is a problem. Copper NiTi overcomes variability in transition temperature and with careful
manipulation of its transition temperature can provide an adequately stiff wire.

We also think it is sensible to match archwire sequences to the patient’s response to initial archwires. This generally
means that older patients have a slower and more progressive archwire progression.

Archwire sequences can be classified in different ways; the preferred method is by archwire type. The four archwire
phases are:

• phase 1
o light round CuNiTi archwires
• phase 2
o rectangular CuNiTi archwires
• phase 3
o major mechanics
ARCHWIRES AND ARCHWIRE TECHNOLOGY 195
EXCELLENCE IN ORTHODONTICS 2016

• phase 4
o finishing

Phase 1: Light round CuNiTi archwires

Aim Archwires Appointment Duration


intervals

Normal 10 weeks 10 to 20 weeks


Level 0.014” CuNiTi
Align Alternatives
Correct majority of rotations 0.013” CuNiTi (more crowded cases)
Initiate arch form 0.016” CuNiTi (following 0.013” CuNitTi)

Phase 1 uses high technology archwires to level, align, almost fully correct all anterior rotations and partially correct
posterior rotations. It also begins to initiate arch form. Normally this phase lasts 10 weeks; however, very crowded
and irregular cases require 20 weeks with 10 weeks in 0.013” CuNiTi and 10 weeks in 0.016” CuNiTi. There is no rush
to get into larger dimension wires – almost all major tooth alignment can be achieved with these archwires.

Phase 2: Rectangular CuNiTi archwires


Aim Archwires Appointment Duration
intervals

Normal 20 to 30 weeks
Fully resolve all rotations 0.014” x 0.025” CuNiTi 8-10 weeks
Begin torque control 0.018” x 0.025” CuNiTi
Anterior space consolidation Alternatives 4-6 weeks
Continue arch form 0.016” x 0.025”CuNiTi (cases with minimal
development irregularity)
Intrusion and bite opening
0.017” x 0.025” RCOS CuNiTi
0.019” x 0.025” RCOS CuNiTi
Torque
0.019” x 0.025” CuNiTi + 20°

Phase 2 fully resolves all rotations; posterior rotations are slowest and most difficult to correct, particularly if they
are the terminal tooth on the archwire. Phase 2 also is the beginning of torque control. It is important to use an
archwire that is 0.025" in the first order dimension to ensure first order alignment of the teeth before progressing to
working archwires if using self-ligating brackets; otherwise the change into the major stainless steel working
archwires will be difficult because of lack of first order alignment.
The 0.016" x 0.025" CuNiTi archwire is used in cases (normally class II division I) with very mild irregularity instead of
the 0.014” x 0.025” and 0.014” x 0.025” CuNiTi archwires.

Intrusion and bite opening is also largely completed at this stage using nickel titanium archwires with reverse curves
of Spee. Additional torque can be achieved using wires with 20° of torque added.

Phase 3: Rectangular non-CuNiTi archwires


Preposted stainless 0.019” x 0.025” stainless steel archwires are the preferred archwires for this stage. Preposted
archwires are less expensive than crimpable hooks and never slip. Their disadvantage is that there may be some
softening of the wire at the point where the hook is soldered to the archwire. Depending on the needs of the case,
196 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

TMA archwires can be substituted for stainless steel particularly where the additional rigidity of the stainless steel
wires is not required.

Aim Archwires Appointment Duration


intervals

Normal
Maintain archform integrity 0.019” x 0.025” stainless steel preposted 6-8 weeks 20 weeks
Finish torque control 0.019” x 0.025” TMA 4 weeks if wearing
Consolidate posterior space Alternative elastics

Corrections 0.016” x 0.025” stainless steel preposted and


anteroposterior crossbite correction with cross elastics
buccolingual
vertical

Phase 4: Finishing archwires


The finishing stages of treatment involves the optimisation of final tooth positions. Minor tooth adjustments can
be made in 0.019” x 0.025” stainless steel archwires. TMA is more resilient and this material is better for larger
corrections. Careful archwire bending and the use of posterior-V elastics can be used to develop interdigitation. If
only minor vertical adjustment without wire bending is required, then 0.014” x 0.025” CuNiTi archwires can be used
to improve buccal segment fit.

Aim Archwires Appointment Duration


intervals

Normal
Make remaining adjustments to: 0.019” x 0.025” stainless steel preposted 4-6 weeks 10-20 weeks
in-out Alternative
angulation
inclination 0.019” x 0.025” TMA for minor adjustments of
rotation tooth position

Develop interdigitation 0.014” x 0.025” CuNiTi allows more vertical


settling while maintaining first order alignment

Two different archwire sequences are commonly used. The three wire archwire sequence is used in relatively well
aligned arches and the four wire archwire sequence used in more irregular arches.
The 0.016" x 0.025" nickel titanium archwire is used as a transitional wire between the initial aligning wires and the
working archwires. It does not provide any torque control except on very poorly aligned teeth. It is important to
use an archwire that is 0.025" in the first order dimension to ensure first order alignment of the teeth before
progressing to working archwires; the reduced slot depth of the Damon System bracket and the precise nature of
the closure mechanism make this more critical.

In the initial aligning and transition stages of treatment, archwires should be left in place for ten weeks. There is no
need to see the patient more frequently than this; if you are uncomfortable with this length of time between patient
appointments then see the patient for a review visit at five weeks but do not change the archwire!

Additional archwires are given in the same table and these may be useful on occasion. Additional information is
given in the chapter on Self Ligating Brackets: Theory and Practice.
ARCHWIRES AND ARCHWIRE TECHNOLOGY 197
EXCELLENCE IN ORTHODONTICS 2016

Archwire error
18
16
14
First order error
12
This occurs when an archwire rotates in a first
10

Deviation angle
order direction within a molar tube or an Activa
bracket. The error is given by the formula: 8
6
4
2
0
16-22 17-25 18-25 19-19 19-25 21-25 21-28
Archwire size (inches x 0.001)
where:
Figure 11.9: Torque error for various sizes of rectangular archwire
w is the width of the bracket
d is the depth of the bracket
dim1 is the first order dimension of the archwire.

Second order error


Tip or second order error occurs because of the use of undersized archwires. The error is related to the diameter of
a round wire or the height of a rectangular wire and the bracket width but becomes negligible in working archwires
as shown in Figure 11.8.

The error is given by the formula:

where:

w is the width of the bracket

h is the height of the bracket

dim2 is the second order dimension of the archwire.


8

Third order error 7


Torque or third order error again occurs 6
because of the use of undersized archwires.
5
The error is due to the fact that the archwire
can rotate within the bracket before 4
degrees

applying any force to the bracket. This 3


error is not negligible in working archwires
2
and is of considerable significance in any
case that requires palatal movement of the 1

upper incisor roots. The magnitude of this 0


error is given in Figure 11.9. 2 2.5 3 3.5 4
bracket width (mm)
0.022" 0.020" 0.018" 0.016" 0.014" 0.012"
The error is given by the formula:
Figure 11.8: Tip error for various bracket widths and archwire sizes
198 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

where:

h is the height of the bracket

dim1 is the first order dimension of the archwire

dim2 is the second order dimension of the archwire.

It can be seen that an 0.016" x 0.022" archwire gives minimal third order control. An 0.018" x 0.025" archwire gives
10 degrees of play while a 0.019" x 0.025" archwire only reduces this play by 2 degrees. Full sized archwires are
therefore required if the torque built into the bracket is to be fully expressed.

However, this takes no account of manufacturing tolerances or radiussing of the archwire corners. Brauchli et al
(2012), demonstrate that the actual play is between 15º and 17 in either direction for most bracket types.

However, the actual situation is worse than this as in addition a further 40% of the torque can be lost if the actual
amount of torque transmitted to the bracket is measured (Sebanc et al 1984). Gioka et al (2004) have cited five
variables which may affect the amount of torque loss between archwire and bracket. These are:

• play between archwire and bracket slot


• lack of stiffness of bracket structure or slot
• inadequate archwire stiffness
• incomplete ligation
• manufacturing variability

All of these factors can be controlled by the clinician except for manufacturing variability. The amount of torque
loss can therefore be reduced by:

• specifying a (substantially) higher value prescription than the final tooth inclination target
• using metal or ceramic brackets or composite brackets with metal slots
• using large cross section stainless steel archwires
• using secure ligation. Elastomeric ligatures lose 40% of their force within the first 24 hours in vitro
and due to enzymatic degradation and thermal cycling, the in vivo situation is likely to be worse. This
is a good reason to use passive self-ligation (or wire ligatures).

Interestingly (and perhaps against intuition until thought about!), Sifikakis et al (2012) have shown that the
torqueing moments generated by a 0.0175” x 0.025” stainless steel wire with 15 degrees of torque in an 0.018” slot
are greater than those generated by a 0.019” x 0.025” stainless steel archwire in an 0.022” slot. In addition, they also
found that higher torque was generated by high
torque brackets than by low torque brackets. The
0.017” x 0.025” in 0.019” x 0.025” in
0.018” slot 0.022” slot results of this study are given in Table 11.6.
High torque 14.3 9.3

Low torque 9.6 6.5


Is it possible to accelerate tooth
Table 11.6: Torquing moments in N-mm for archwires in movement?
0.018” and 0.022” slots (from Sifikakis et al 2012)
ARCHWIRES AND ARCHWIRE TECHNOLOGY 199
EXCELLENCE IN ORTHODONTICS 2016

Several methods of accelerating tooth movement have been described and these include:

• corticotomies
• interseptal bone reduction
• low level laser therapy
• pulsed electromagnetic fields
• vibrational therapy
• injected substances (Relaxin)

El-Angbawi et al (2015) conducted a Cochrane systematic review of non-surgical adjunctive interventions for
accelerating tooth movement and found only two papers that met the review criteria; Miles et al 2012 and Pavlin et
al (2015). The authors concluded that there was little clinical research about the effectiveness of non-surgical
interventions to accelerate orthodontic treatment and that the current evidence was of very low quality. Although
there have been claims that there may be a positive effect of light vibrational forces, results of the current studies
do not reach either statistical or clinical significance.

Woodhouse et al (2015) investigated the use of vibrational force in a prospective randomised controlled trial on
tooth alignment and found no evidence that supplemental vibrational force can significantly increase the rate of
initial tooth movement or reduce the amount of time required to achieve final alignment when used in conjunction
with a preadjusted edgewise fixed appliance.

At the moment, there is insufficient evidence to support the use adjunctive vibrational forces to speed up
orthodontic treatment.

Fleming et al (2015) carried out a Cochrane systematic review of surgical adjunctive interventions for accelerating
tooth movement and found that there was limited research concerning the effectiveness of surgical interventions
to accelerate orthodontic treatment. The authors concluded that on short-term outcomes these procedures do
appear to show promise as a means of accelerating tooth movement and that it is possible that these procedures
may prove useful; however, further prospective research comprising assessment of the entirety of treatment with
longer follow-up is required to confirm any possible benefit.

Pain from initial archwires


Again little work has been carried out on discomfort from archwires. Jones (1984) and Jones and Chan (1992) looked
at the discomfort caused by initial aligning archwires (multistrand and superelastic NiTi) using tooth extraction as a
control and found that:

• the pain from archwires was greater than from tooth extraction
• the pain lasted 5-6 days
• the amount of discomfort increased with age
• there was little difference in the discomfort produced by either archwire

This contrast with the findings of Ngan (1989) who found no difference in discomfort between patients under and
over 16 years of age. This may be explained by the fact that Jones used a verbal scale to determine pain levels
compared with Ngan’s visual analogue scale (VAS).

Ngan (1994) has compared the effectiveness of a single dose of a placebo, aspirin and ibuprofen in reducing pain
after the placement of separators and initial orthodontic archwires. Both aspirin (650 mg) and ibuprofen (400 mg)
reduced discomfort with ibuprofen being significantly more effective than aspirin.
200 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Leavitt et al (2002) have published an interesting paper which looks at pulpal pain during orthodontic treatment –
in the past it has been assumed that the pain associated with tooth movement was periodontal pain. This evaluated
detection and pain thresholds in the dental pulp and compared them with a pain visual analogue scale (VAS)
completed by the patient. There was a significant difference between the experimental and control groups at post-
archwire insertion day 1 followed by a small but non-significant difference at post-archwire insertion day 7. Despite
finding a non-significant but suggestive (there was an outlier in the data) negative correlation between pain scores,
the investigators were unable to find significant differences in the detection or pain threshold at any of the
assessment periods. This paper is strongly suggestive of the fact that pulpal pain is involved in the pain from
orthodontic appliances but does not quite manage to show it!

Erdinç and Dincer (2004) showed that pain was initiated approximately two hours after initial archwire placement,
that 90% of patients experienced pain in the first week and that the mean pain intensity peaked at 24 hours and
thereafter gradually declined over 7 days.

Bartlett et al (2005) demonstrated that a telephone call from the orthodontic provider reduced patient’s self-
reported pain and anxiety scores. This was a well-designed and interesting study having two experimental groups
and one control group which were asked to complete pain intensity and anxiety state questionnaires in the week
following appliance placement. The first experimental group received a structured telephone call within 24 hours
of appliance placement covering:

• the patient’s well being


• the pain and discomfort experienced
• reassurance that the patient’s discomfort was within normal limits
• the need to maintain excellent oral hygiene
• the need for a soft diet
• a reminder to use analgesics if necessary
• and the importance of maintaining a positive attitude

The second experimental group also received a similarly timed telephone call but this simply thanked the patient
for agreeing to participate in the study and reminding them to fill in the questionnaire. The control group received
no telephone call.

Both groups receiving a telephone call reported lower pain intensity scores and lower anxiety states than the control
group – the content of the telephone call was not important.

Pringle et al (2009) have demonstrated that the mean pain intensity in the first week after initial archwire placement
is lower with the Damon passive self-ligating appliance than with conventional appliances.

In a recent paper, Abdelrahman et al (2015b) investigated whether there was a difference in pain perception in
patients being given different types of initial archwire.

The archwires were:

• 3M Unitek conventional NiTi


• 3M Unitek superelastic NiTi
• 3M Unitek thermoelastic NiTi

The various groups were matched for age, gender, degree of initial crowding, malocclusion (incisor classification),
and type of treatment (extraction vs nonextraction). Pain perception was recorded using a ten-point visual
ARCHWIRES AND ARCHWIRE TECHNOLOGY 201
EXCELLENCE IN ORTHODONTICS 2016

analogue scale. No significant difference was found in pain intensity between the three types of archwire was
experienced by patients during initial tooth alignment. In addition, gender, age, and the degree of crowding have
no effect on the perceived discomfort experienced by patients undergoing fixed orthodontic treatment.

Tooth movements
Management
Tooth movement is often described as first, second and third order; however, this can be misleading. Tooth
movement can occur in each of three directions at right angles to one another – the x, y and z axes. In addition,
these movements can occur as bodily movement or a moment about the axis. These movements are shown in
Figures 11.10 and 11.11.

Basic wire bending skills


Although the Straight-Wire Appliance greatly reduces the amount of wire bending required during active
treatment, orthodontists using the appliance still require basic wire bending skills. These should include:

• an understanding of first, second and third order bends


• the ability to bend flat, symmetrical and co-ordinated archwires
• the ability to bend and adjust bayonet bends and various designs of loop
• the ability to incorporate and adjust progressive and continuous torque
• the ability to bend and adjust curves of Spee

Ligation
Proper ligation technique is as important as accurate band and bracket positioning.

Wire ligature ties


Operators should be familiar with the use of conventional ties, rotation ties and lacebacks tied with 0.010" wire
ligatures and Coon pattern ligature lockers. In practice however these are now rarely used as the use of elastomeric
ligatures is preferred in conventional ligation.

Figure 11.10: Translational movement in the x, y Figure 11.11: Moments around the x, y and z axes (pitch, roll
and z axes corresponds to mesiodistal, up and down and in-out and yaw) correspond to angulation, inclination and derotation of teeth
movements of teeth
202 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

3.5
Elastomeric ties
3 Elastomeric ligatures have for many years been used
routinely as the predominant method of ligation. These
2.5
are best placed with Orthopli 018R or Microna forceps and
2 are used in a figure of eight pattern where additional force
ratio

1.5 is required to obtain full bracket engagement. Figure of


eight elastomeric ligatures should be used selectively to
1
fine tune bracket engagement and resistance to sliding.
0.5 The use of figure of eight ligatures increases the fictional
0 resistance by approximately one and a half times for most
016 x 022 017 x 025 018 x 025 019 x 025 working archwires and by over three times for 0.016" x
Figure 11.12: Comparison of ratio of frictional resistance
0.022" archwires (Sims et al 1993).
between figure of eight and conventional elastomeric ligation (from
Sims et al 1993) Hain et al (2006) studied the effects of different types of
ligation on sliding friction with an 0.0119” x 0.025”
stainless steel archwire. The ligation methods tested included Damon 2 and SPEED self-ligating brackets, TP
SuperSlick, silicone impregnated and conventional elastomeric ties. The Damon 2 self-ligating brackets
demonstrated the least friction followed by the TP SuperSlick elastomeric ligatures whose coating seemed resistant
to wear in a simulated clinical setting.

Kobayashi ligatures
Kobayashi ligatures are used to add hooks to brackets. These ligatures should be reasonably soft but all proprietary
Kobayashi ligatures are unsuitable for use on ceramic brackets - it is better to make up a hook from an 0.010" wire
tie instead for these cases. The use of Kobayashi ligatures is declining.

Archform
Archform is usually interpreted as meaning the two dimensional shape of the dental arch viewed from the occlusal;
a little further thought would elicit the comment that size was also important.

The development of pre-adjusted edgewise appliances has stimulated the wish for preformed archwires which
simulate “ideal” archform to be commercially available. Many manufacturers offer archwires in different sizes
(widths) and some in both different widths and shapes.

Much has been written on the shape and size of the ideal natural dental arch and the derivation of archforms for
orthodontic use. Archform has generally been described using:

• simple mathematical descriptions


o segments of a circle linked to straight lines (Hawley 1905)
o ellipses (Brader 1972)
o parabolas (Jones and Richmond 1989)
o catenary curve (Pepe 1975)
• complex mathematical curve-fitting formulae
o conic sections (Sampson 1981)
o cubic spline (BeGole 1979)
o polynomial function (Lu 1966)
o “mixed” functions (Ferrario et al 1994)
o beta function (Braun et al 1998)
ARCHWIRES AND ARCHWIRE TECHNOLOGY 203
EXCELLENCE IN ORTHODONTICS 2016

Knox et al (1994) provide a thought-provoking look at the possibility of an ideal preformed archwire. The notion of
an ideal orthodontic archform, with its implications of symmetry and mathematical description, does not accord
with what is found in nature. In particular, symmetry is not a tenable assumption. Furthermore, the development
of the preadjusted edgewise appliances means that archwires no longer follow the contour of the buccal cusps and
incisal edges. As manufacturers make more sophisticated appliances, archform may well deviate further from the
smooth archform described in an “ideal” arch (see Orthos archform where CIS creates a sharp bend in the canine
region). When comparing arches used with the Straight-Wire Appliance with edgewise arches, it is also important
to remember that to produce core arch lines of the same shape and size, a preadjusted edgewise appliance arch
must be wider in the premolar region. This reflects the built in molar offset.

The shape of dental arch that we should adopt as our treatment goal has long been a topic for lively discussion.
Particular archforms have been advocated on grounds of producing better intercuspal occlusion, better functional
occlusion, better stability and better aesthetics.

Because differences in arch shape are so noticeable to the eye, it is important to have a more accurate idea of the
extent to which differences in arch shape produce differences in arch dimension. Superimposing arches of different
shape reveals that arch widths vary by less than might be guessed and that the differences are frequently in the
premolar and not the canine region. Measurements on typodonts quantify the effects of various shapes on
intercanine, interpremolar and intermolar widths.

The paper by Felton in 1987 investigated different archforms with regard to the natural occurrence of the various
shapes in untreated malocclusions and the stability of the change of shape produced by the different archforms
during treatment. No archform was the closest match for more than 20% of cases. Regarding stability, of the 65%
that had a change of shape during treatment, 70% returned to the pre-treatment shape. Felton did not report
whether this was associated with a return of dental irregularity, but the paper did support the concept of not
expanding intercanine width and suggests that some customising of arch shape may produce more stable results.

Leaving aside archform, there is much evidence over many years to support the practice of not significantly altering
arch widths during treatment, although almost all orthodontists do! Huntley (1990) has provided interesting
additional evidence in his study of cases all treated with arches of the same size. Approximately half these cases
had a significant increase in intercanine width. If preformed arches are left unaltered in size, they should therefore
be purchased in different sizes.

Ricketts has developed an approach that combines variability in shape with variation in size. He felt that larger
arches also tend to be of a different shape. A combination of three different shapes and smaller sizes of two of the
shapes therefore comprise a “Pentamorphic” system. This is based on studies of ideal and stable treated occlusions
and in particular on one study of only 40 patients. The basis for choosing an arch for a patient is not very clear. A
“rough estimate can be made from gross facial form”, but analysis of lateral and frontal cephalometric radiographs
by the Rocky Mountain Diagnostic Services is advocated as an “immense advantage”.

Robnett (1980) said that “A universal archform probably does not exist. Years of seeking such a form have resulted
in limited improvements in orthodontic care, and yet the ideal form remains elusive.” Customising archform is
therefore essential.

However, CBCT images show that the teeth are frequently not positioned over the apical bases but inclined
lingually. This raises questions about whether in fact the arches can be expanded to reposition the teeth over the
basal bone or whether the original arch shape reflects the size of underlying dent-alveolar arches. Ronay et al (2008)
have investigated the relationship between the dental and basal archform using the FA points to represent the
dental arch and the WALA points to represent the basal arch. The WALA points were described by Andrews and
Andrews (2000) described the WALA ridge as the most prominent point on the soft-tissue ridge immediately
204 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

occlusal to the mucogingival junction. It is located at 40


or nearly at the same vertical level as the horizontal
centre of rotation of each tooth.

Ronay et al (2008) found that the arch form derived 20


from the WALA points was much broader in the

mm
premolar and molar regions and that all archforms
were highly individual and could not be generalised
by a single shape. There was significant correlation 0
between the FA and WALA points particularly in the -40 -20 0 20 40
canine and molar regions and that the WALA points mm
WALA points FA points
could be used to indicate basal archform.
Figure 11.13: Superimposition of the FA point and WALA curves
A questionnaire study carried out at Bristol Dental for the lower arch (after Ronay et al 2008)
Hospital (McNamara et al 2009) to determine the
choices made by clinicians with respect to archwires and arch form during the early and late stages of treatment
with fixed appliances. Most clinicians felt that maintenance of the pretreatment arch form, particularly intercanine
width, was important in the later stages of treatment although it was not considered so important in the early stages
of treatment when NiTi archwires were used. There was however no consistency as to how arch form should be
measured or determined, and methodologies were used in different ways by different clinicians.

McNamara et al (2010) further reported that comparing the intercanine and intermolar width measurements on
archwires fabricated on three sets of standardized typodonts models demonstrated differences in the mean values
of each dimension, even though there were no differences on the models. In treated patients, large variations in
intercanine and intermolar widths were also observed for both intraoperator and interoperator figures. Some
clinicians contracted the intercanine and intermolar widths, others expanded them. Although most clinicians aimed
to maintain pretreatment arch forms, this study showed that this was not often transferred to clinical practice

Tooth 1 2 3 4 5 6

Mean -1.21 -0.88 -0.32 0.59 1.78 2.77

sd 1.24 1.07 1.63 1.28 1.10 0.89

Table 11.7: The mean difference and standard deviations between corresponding FA and WALA points in the mandible (n=70) (from
Ronay et al 2008)

Archform and arch width


Lee (1999) has given a good review of archform and arch width. This gives a good summary of conventional thought
on the possibilities for obtaining stable arch expansion. It emphasises that arch expansion is likely to be more stable
in the absence of extractions, the immutability of lower canine width, correction of class II arch relationships may
result in some stable expansion and that arch expansion produces space of only 33% of the expansion in the
posterior region.

References
Abdelrahman RS, Al-Nimri KS and Al Masitah EF (2015)
A clinical comparison of three aligning archwires in terms of alignment efficiency: A prospective clinical trial
Angle Orthodontist 85: 434-439

Abdelrahman RS, Al-Nimri KS and Al Masitah EF (2015b)


Pain experience during initial alignment with three types of nickel-titanium archwires: A prospective clinical trial
Angle Orthodontist 85: 1021-1026
ARCHWIRES AND ARCHWIRE TECHNOLOGY 205
EXCELLENCE IN ORTHODONTICS 2016

Andrews LF and Andrews WA (2000)


The six elements of orofacial harmony
Andrews Journal 1: 13-22

Badran SA, Orr JF, Stevenson M and Burden DJ (2003)


Photo-elastic stress analysis of initial alignment archwires
European Journal of Orthodontics 25: 117-125

Bartlett BW, Firestone AR, Vig KWL, Beck M and Marucha PT (2005)
The influence of a structured telephone call on orthodontic pain and anxiety
American Journal of Orthodontics and Dentofacial Orthopaedics 128: 435-441

BeGole EA (1980)
Application of cubic spline function in the description of dental arch form
Journal of Dental Research 59: 1549-1556

Berger JL (2008)
The SPEED System: an overview of the appliance and clinical performance
Seminars in Orthodontics 14: 54-62

Bishara SE, Winterbottom JM, Suleiman A-HA, Rim K and Jakobsen JR (1995)
Comparisons of thermodynamic properties of three nickel titanium orthodontic archwires
Angle Orthodontist 65: 117-122

Bolender Y, Vernière A, Rapin C and Filleul M-P (2010)


Torsional superelasticity of NiTi archwires: myth or reality?
Angle Orthodontist 80: 1100-1109

Brader AC (1972)
Dental arch form related with intraoral forces: PR = C
American Journal of Orthodontics 61: 541-61

Bradley TG, Brantley WA and Culbertson BM (1996)


Differential scanning calorimetry (DSC) analysis of superelastic and nonsuperelastic nickel-titanium orthodontic wires
American Journal of Orthodontics and Dentofacial Orthopaedics 109: 589-597

Brantley WA, Iijima M and Grentzer TH (2003)


Temperature-modulated DSC provides new insight about nickel titanium wire transformations
American Journal of Orthodontics and Dentofacial Orthopaedics 124: 387-394

Brauchli LM, Steineck M and Wichelhaus A (2012)


Active and passive self-ligation: a myth? Part 1: torque control
Angle Orthodontist 82: 663-669

Braun S, Hnat WP, Fender DE and Legan HL (1998)


The form of the human dental arch
Angle Orthodontist 68: 29-36

El-Angbawi A, McIntyre GT, Fleming PS and Bearn DR (2015)


Non-surgical adjunctive interventions for accelerating tooth movement in patients undergoing fixed orthodontic
treatment
Cochrane Database Systematic Review 2015 Nov 18;11:CD010887. doi: 10.1002/14651858.CD010887.pub2

Eliades T and Bourauel C (2005)


Intraoral aging of orthodontic materials: the picture we miss and its clinical relevance
American Journal of Orthodontics and Dentofacial Orthopaedics 127: 403-12

Erdinç AM and Dincer B (2004)


Perception of pain during orthodontic treatment with fixed appliances
European Journal of Orthodontics 26: 79-85

Evans TJW and Durning P (1996a)


Orthodontic Products Update
British Journal of Orthodontics 23: 1-4

Evans TJW and Durning P (1996b)


Orthodontic Products Update
Aligning archwires, the shape of things to come? - A fourth and fifth phase of delivery
British Journal of Orthodontics 23: 269-275
206 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Evans TJW, Jones ML and Newcombe RG (1998)


Clinical comparison and performance perspective of three aligning archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 114: 32-39

Felton JM, Sinclair PM, Jones DL and Alexander RG (1987)


A computerised analysis of the shape and stability of mandibular archform
American Journal of Orthodontics and Dentofacial Orthopaedics 92: 478-583

Ferrario VF, Sforza C, Miani A and Tartaglia G (1994)


Mathematical definition of the shape of dental arches in the human permanent healthy dentitions
European Journal of Orthodontics 16: 287-294

Filleul MP and Jordan l (1997)


Torsional properties of Ni-Ti and Copper-Ni-Ti wires: the effect of temperature on physical properties
European Journal of Orthodontics 19: 637-646

Fleming PS, Fedorowicz Z, Johal A, El-Angbawi A and Pandis N (2015)


Surgical adjunctive procedures for accelerating orthodontic treatment
Cochrane Database of Systematic Reviews 2015, Issue 6. Art. No.: CD010572. DOI: 10.1002/14651858.CD010572.pub2

Gatto E, Matarese G, Di Bella G, Nucera R, Borsellino C and Cordasco G (2011)


Load-deflection characteristics of superelastic and thermal nickel-titanium archwires
European Journal of Orthodontics first published online October 23, 2011 doi:10.1093/ejo/cjr103
Accessed 23 January 2012

Gioka C and Eliades T (2004)


Materials-induced variation in the torque expression of preadjusted appliances
American Journal of Orthodontics and Dentofacial Orthopaedics 125: 332-338

Gurgel J de A, Herr S, Powers JM and LeCrone V (2001)


Force-deflection characteristics of superelastic nickel titanium archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 120: 378-382

Hain M, Dhopatkar A and Rock P (2006)


A comparison of different ligation methods on friction
American Journal Orthodontics Dentofacial Orthopaedics 130: 666-670

Hawley CA (1905)
Determination of the normal arch and its application to orthodontics
Dental Cosmos 47: 541-557

Hilgers JJ (1999)
Reverse curve TMA with T-loops
http://www.ormco.com/index/ormco-education-video-reverse-curve-tma-with-t-loops
Accessed: 12 February 2011

Hurst CL, Duncanson MG, Nanda RS and Angolkar PV (1990)


An evaluation of the shape-memory phenomenon of nickel-titanium orthodontic wires
American Journal of Orthodontics and Dentofacial Orthopaedics 98: 72-76

Huntley P (1990)
The effects of preformed ‘ideal’ archwires upon arch dimensions
British Journal of Orthodontics Research Report 16: 64

Jian F, Lai W, Furness S, McIntyre GT, Millett DT, Hickman J, Wang Y (2013)
Initial arch wires for tooth alignment during orthodontic treatment with fixed appliances
Cochrane Database of Systematic Reviews 2013, Issue 4. Art. No.: CD007859. DOI: 10.1002/14651858.CD007859.pub3

Jones ML (1984)
An investigation into the initial discomfort caused by placement of an archwire
European Journal of Orthodontics 6: 48-54

Jones ML and Richmond S (1989)


An assessment of the fit of a parabolic curve to pre- and post-treatment dental arches
British Journal of Orthodontics 16: 85-93
ARCHWIRES AND ARCHWIRE TECHNOLOGY 207
EXCELLENCE IN ORTHODONTICS 2016

Jones ML and Chan C (1992)


The pain and discomfort experienced during orthodontic treatment: a randomised controlled trial of two initial
aligning archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 102: 373-381

Khier SE, Brantley WA and Fournelle RA (1991)


Bending properties of superelastic and nonsuperelastic nickel-titanium orthodontic wires
American Journal of Orthodontics and Dentofacial Orthopaedics 99: 310-318

Knox J, Jones ML and Durning P (1994)


An ideal preformed archwire?
British Journal of Orthodontics 20: 65-70

Kula K, Phillips C, Gibilaro A and Proffit WR (1998)


Effect of ion implantation of TMA archwires on the rate of orthodontic sliding space closure
American Journal of Orthodontics and Dentofacial Orthopaedics 114:577-580

Kusy RP (1991)
Nitinol alloys: so, who's on first? (Letter to the editor)
American Journal of Orthodontics and Dentofacial Orthopaedics 100: 25A

Kusy RP (1997)
A review of contemporary archwires; their properties and characteristics
Angle Orthodontist 67: 197-207

Laheurte P, Eberhardt A, Philippe MJ and Deblock L (2007)


Improvement of pseudoelasticity and ductility of Beta III titanium alloy—application to orthodontic wires
European Journal of Orthodontics 29: 8-13 (Open access article)

Leavitt AH, King GJ, Ramsay DS and Jackson DL (2002)


A longitudinal evaluation of pulpal pain during orthodontic tooth movement
Orthodontics and Craniofacial Research 5: 29-37

Lee RT (1999)
Arch width and form: a review
American Journal of Orthodontics and Dentofacial Orthopaedics 115: 305-313

Lu KH (1966)
An orthogonal analysis of the form, symmetry and asymmetry of the dental arch
Archives of Oral Biology 11: 1057-1069

Mandall NA, Lowe C, Worthington HV, Sandler J, Derwent S, Abdi-Oskouei M and Ward S (2006)
Which orthodontic archwire sequence? A randomized clinical trial
European Journal of Orthodontics 28: 561–566

Matasa CG (1997)
NiTi alloys; two metals in one
The Orthodontic Materials Insider 10: 1; 2-7

McNamara C, Drage KJ, Sandy JR and Ireland AJ (2009)


An evaluation of clinicians’ choices when selecting archwires
European Journal of Orthodontics 32: 54-59

McNamara C, Sandy JR and Ireland AJ (2010)


Effect of arch form on the fabrication of working archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 138: 257.e1-257.e8

Meling TR and Ødegaard J (1998a)


The effect of temperature on the elastic responses to longitudinal torsion of rectangular nickel-titanium archwires
Angle Orthodontist 68: 357-368

Meling TR and Ødegaard J (1998b)


The effect of short-term temperature changes on the mechanical properties of rectangular nickel-titanium archwires
tested in torsion
Angle Orthodontist 68: 369-376
208 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Meling TR and Ødegaard J (1998c)


Short-term temperature changes influence the force exerted by superelastic nickel-titanium archwires activated in
orthodontic bending
American Journal of Orthodontics and Dentofacial Orthopaedics 114: 503-509

Miles P, Smith H, Weyant R and Rinchuse DJ (2012)


The effects of a vibrational appliance on tooth movement and patient discomfort: a prospective randomised clinical
trial
Australian Orthodontic Journal 28: 213-218

Mittal N, Xia Z, Chen J, Stewart KT and Liu SSY (2012)


Three-dimensional quantification of pretorqued nickel-titanium wires in edgewise and prescription brackets
The Angle Orthodontist (in press) pp. (not available) doi: http://dx.doi.org/10.2319/062812-532.1

Miura F, Masakuni M, Ohura Y and Hamanaka H (1986)


The super-elastic property of the Japanese NiTi alloy wire for use in orthodontics
American Journal of Orthodontics and Dentofacial Orthopaedics 90: 1-10

Mizrahi E, Cleaton-Jones PE, Luyckx S and Fatti LP (1991)


The effect of Ion implantation on the beaks of orthodontic pliers
American Journal of Orthodontics and Dentofacial Orthopaedics 99: 513-519

Nakano H, Satoh K, Norris R, Jin T, Kamegai T, Ishikawa F and Katsura H (1999)


Mechanical properties of several nickel-titanium alloy wires in three-point bending tests
American Journal of Orthodontics and Dentofacial Orthopaedics 115: 390-395

Ngan P, Kess B and Wilson S (1989)


Perception of discomfort by patients undergoing orthodontic treatment
American Journal of Orthodontics and Dentofacial Orthopaedics 96: 47-53

Ngan P, Wilson S, Shanfield J and Amini H (1994)


The effect of ibuprofen on the level of discomfort in patients undergoing orthodontic treatment
American Journal of Orthodontics and Dentofacial Orthopaedics 106: 88-95

O'Brien K, Lewis D, Shaw W and Coombe E (1990)


A clinical trial of aligning archwires
European Journal of Orthodontics 12: 380-384

Pandis N, Polychronopoulou A and Eliades T (2009)


Alleviation of mandibular anterior crowding with copper-nickel-titanium vs nickel-titanium wires: A double-blind
randomized control trial
American Journal of Orthodontics and Dentofacial Orthopaedics 136: 152.e1-152.e7

Pavlin D, Anthony R, Raj V and Gakunga PT (2015)


Cyclic loading (vibration) accelerates tooth movement in orthodontic patients: A double-blind, randomized
controlled trial
Seminars in Orthodontics 21:187-194

Pepe SH (1975)
Polynomial and catenary curve fits to human dental arches
Journal of Dental Research 54: 1124-1132

Pringle AM, Petrie A, Cunningham SJ and McKnight M (2009)


Prospective randomised clinical trial to compare pain levels associated with two orthodontic fixed bracket systems
American Journal of Orthodontics and Dentofacial Orthopaedics 36: 160-167

Proffit WR (1986)
Contemporary Orthodontics
St Louis, The CV Mosby Company

Ren C, Bai Y, Wang H, Zheng Y and Li S (2008)


Phase transformation analysis of varied nickel-titanium orthodontic wires
Chinese Medical Journal 121: 2060-2064

Ricketts RM (1979)
Design of archform and details for bracket placement
Distributed by Rocky Mountain Orthodontics
ARCHWIRES AND ARCHWIRE TECHNOLOGY 209
EXCELLENCE IN ORTHODONTICS 2016

Riley M and Bearn D (2008)


A systematic review of clinical trials of aligning archwires
Journal of Orthodontics 36: 42-51

Robnett JH (1980)
Segment concept in arch pattern design
American Journal of Orthodontics and Dentofacial Orthopaedics 77: 355-367

Ronay V, Miner RM, Will LA and Arai K (2008)


Mandibular arch form: The relationship between dental and basal anatomy
American Journal of Orthodontics and Dentofacial Orthopaedics 
1 34: 430
-438

Rucker BK and Kusy RP (2002a)


Elastic flexural properties of multistranded stainless steel versus conventional nickel titanium archwires
Angle Orthodontist 72: 302-309

Rucker BK and Kusy RP (2002b)


Resistance to sliding of stainless steel multistranded archwires and comparison with single-stranded levelling wires
American Journal of Orthodontics and Dentofacial Orthopaedics 122: 73-83

Rucker BK and Kusy RP (2002c)


Elastic properties of alternative versus single stranded levelling archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 122: 528-541

Ryan R, Walker G, Freeman K and Cisneros GJ (1997)


The effects of ion implantation on rate of tooth movement: an in vitro model
American Journal of Orthodontics and Dentofacial Orthopaedics 112: 64-68

Sampson PD (1981)
Dental arch shape: a statistical analysis of arch shape using conic sections
American Journal of Orthodontics 79: 535-548

Santoro M, Nicolay OF and Cangialosi TJ (2001a)


Pseudoelasticity and thermoelasticity of nickel titanium alloys: a clinically oriented review. Part 1: Temperature
transitional ranges
American Journal of Orthodontics and Dentofacial Orthopaedics 119: 587-593

Santoro M, Nicolay OF and Cangialosi TJ (2001b)


Pseudoelasticity and thermoelasticity of nickel titanium alloys: a clinically oriented review. Part 1: Deactivation forces
American Journal of Orthodontics and Dentofacial Orthopaedics 119: 594-603

Santoro M and Beshers DB (2000)


Nickel-titanium alloys: stress related temperature transitional range
American Journal of Orthodontics and Dentofacial Orthopaedics 118: 685-692

Sebanc J, Brantley WA, Pincsak JJ, and Conover JP (1984)


Variability of effective root torque as a function of edge bevel on orthodontic arch wires
American Journal of Orthodontics 86: 43-51

Segner D and Ibe D (1995)


Properties of superelastic wires and their relevance to orthodontic treatment
European Journal of Orthodontics 17: 395-402

Sifakakis I, Pandis N, Makou M, Eliades T, Katsaros C and Bourauel C (2012)


Torque expression of 0.018 and 0.022 inch conventional brackets
European Journal of Orthodontics first published online July 24, 2012 doi:10.1093/ejo/cjs041

Sims A P T, Waters N E, Birnie D J and Pethybridge R J (1993)


A comparison of the forces required to produce tooth movement in vitro using two self-ligating brackets and a pre-
adjusted bracket employing two types of ligation
European Journal of Orthodontics 15: 377-385

Tonner RIM and Waters NE (1994)


The characteristics of superelastic Ni-Ti wires in three-point bending. Part 1: The effect of temperature
The characteristics of superelastic Ni-Ti wires in three-point bending. Part 2: intra-batch variation
European Journal of Orthodontics: 16: 409-419 and 421-425
210 ARCHWIRES AND ARCHWIRE TECHNOLOGY
EXCELLENCE IN ORTHODONTICS 2016

Verstrynge A, Van Humbeeck J and Willems G (2006)


In-vitro evaluation of the material characteristics of stainless steel and beta-titanium orthodontic wires
American Journal of Orthodontics and Dentofacial Orthopaedics 130: 460-470

Waters NE (1992)
Superelastic nickel titanium wires
British Journal of Orthodontics 19: 319-322

West AE, Jones ML and Newcombe RG (1995)


Multiflex versus superelastic: a randomised clinical trial of the tooth alignment ability of initial archwires
American Journal of Orthodontics and Dentofacial Orthopaedics 108: 464-471

Weilland F (2003)
Constant versus dissipating forces in orthodontics: the effect on initial tooth movement and root resorption
European Journal of Orthodontics 25: 335-342

Woodhouse NR, DiBiase AT, Johnson N, Slipper C, Grant J, Alsaleh M, Donaldson ANA and Cobourne MT (2015)
Supplemental vibrational force during orthodontic alignment: a randomized trial
Journal of Dental Research 94: 682-689

Zinelis S, Eliades T, Pandis N, Eliades G, Bourauel C (2007)


Why do nickel-titanium archwires fracture intraorally? Fractographic analysis and failure mechanism of in-vivo
fractured wires
American Journal of Orthodontics and Dentofacial Orthopaedics 132: 84-89

Useful related references not referred to in this chapter


Lu KH (1964)
Analysis of dental arch symmetry (abstract)
Journal of Dental Research 43: 780

You might also like