You are on page 1of 8

Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Contents lists available at ScienceDirect

Journal of the Mechanical Behavior of


Biomedical Materials
journal homepage: www.elsevier.com/locate/jmbbm

Force delivery of NiTi orthodontic arch wire at different magnitude of MARK


deflections and temperatures: A finite element study

M.F. Razalia, A.S. Mahmuda, , N. Mokhtarb
a
Nanofabrication and Functional Materials Research Group, School of Mechanical Engineering, Engineering Campus, Universiti Sains Malaysia, 14300 Penang, Malaysia
b
Craniofacial and Biomaterial Science Cluster, Advanced Medical and Dental Institute, Universiti Sains Malaysia, Bandar Putra Bertam, 13200 Kepala Batas, Penang,
Malaysia

A R T I C L E I N F O A B S T R A C T

Keywords: NiTi arch wires are used widely in orthodontic treatment due to its superelastic and biocompatibility properties.
Bending In brackets configuration, the force released from the arch wire is influenced by the sliding resistances developed
Leveling on the arch wire-bracket contact. This study investigated the evolution of the forces released by a rectangular
Binding NiTi arch wire towards possible intraoral temperature and deflection changes. A three dimensional finite ele-
Bracket friction
ment model was developed to measure the force-deflection behavior of superelastic arch wire. Finite element
Shape memory alloy
analysis was used to distinguish the martensite fraction and phase state of arch wire microstructure in relation to
the magnitude of wire deflection. The predicted tensile and bending results from the numerical model showed a
good agreement with the experimental results. As contact developed between the wire and bracket, binding
influenced the force-deflection curve by changing the martensitic transformation plateau into a slope. The arch
wire recovered from greater magnitude of deflection released lower force than one recovered from smaller
deflection. In contrast, it was observed that the plateau slope increased from 0.66 N/mm to 1.1 N/mm when the
temperature was increased from 26 °C to 46 °C.

1. Introduction properties of superelasticity, which permits recovery of very large strain


of approximately 6–8% without permanent deformation (Otsuka and
Orthodontic treatments involve the movement of tooth in various Wayman, 1999). This superelastic behavior is found to be suitable for
directions and orientations at different stages. These stages are divided aligning and leveling treatment, in which it provides a constant mag-
into alignment and leveling, correction of molar relationship and space nitude of force for effective tooth movement. As the arch wire slides to
closure, and finishing treatment. Movement of tooth is induced by force recover its original straight shape, it pulls the teeth to a better position
generated from a pre-deformed wire hooked to the bracket on the tooth. (Singh, 2015). It is well documented that superelastic NiTi material
The design of bracket is supposed to allow the arch wire to slide at yields this large strain over a stress plateau by manifesting its stress-
negligible friction, thus tooth movement is achieved only by the induced martensitic (SIM) transformation (Jiang et al., 2009).
bending force generated by the pre-deformed wire. Most brackets are Even though NiTi arch wires are widely used, its real performance in
made of stainless steel and arch wire is secured onto it by using ligation bracket configuration is still unclear since most of the force-deflection
ties. For alignment and leveling stage, orthodontist starts with round studies focused on three point bending test (Kaphoor and
light wire that can be bent easily and subsequently replaces it with a Sundareswaran, 2012; Varela et al., 2014). The force-deflection beha-
thicker wire with rectangular shape (Kim et al., 2008). This sequence of vior deduced from three point bending test is not directly transferable
wires usage is important to gain light forces throughout the orthodontic to the oral clinical setting due to the absence of sliding resistance
treatment, as tooth moves effectively at force below 1.0 N (Proffit et al., consideration during the test. In order to obtain realistic clinical force,
2014). Nucera et al. (2014) have included bracket engagement in their bending
For decades, the use of superelastic nickel-titanium (NiTi) wire is test to consider the influence of sliding resistance towards the bending
preferred against stainless steel wire during the early stage of ortho- behavior of NiTi arch wire. They found that the force released by the
dontic treatment, with a goal of bringing the teeth into alignment 0.36 mm round NiTi arch wire during recovery increased linearly from
(Duerig et al., 2013). NiTi arch wires are chosen due to its unique 3 mm to 1 mm, which completely defies the believe of constant and


Corresponding author.
E-mail address: abdus@usm.my (A.S. Mahmud).

http://dx.doi.org/10.1016/j.jmbbm.2017.09.021
Received 10 June 2016; Received in revised form 20 May 2017; Accepted 14 September 2017
Available online 18 September 2017
1751-6161/ © 2017 Elsevier Ltd. All rights reserved.
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

continuous force delivery of NiTi arch wire in orthodontic treatment.


The sliding resistance in arch wire-bracket configuration was cate-
gorized by Kusy and Whitley (1999) into two components, classic
friction (FL) and binding (FBI). Classic friction refers to sliding resistance
created by elastomer ligatures when it drives the arch wire against the
base of the bracket slot. Meanwhile, binding is developed when the arch
wire is bent, with the magnitude of friction increases as the curvature of
the bend increases. Since arch wire slides through the slot of the ad-
jacent brackets during leveling treatment, a portion of the released
force (FR) from the NiTi arch wire is used up to overcome the generated
binding and classical friction at the wire-bracket interface (Baccetti
et al., 2009).
Researchers have developed various constitutive models of shape
memory alloy using different state variables to describe the phase
transition characteristics such as, reorientation of martensite, marten-
site volume fraction and energy dissipation. A few macro scales con-
stitutive models of shape memory alloy have been proposed in litera-
ture but feature only the superelastic (Auricchio and Taylor, 1997) or
Fig. 1. NiTi arch wire bent on modified three point bending setup.
both shape memory effect and superelastic behavior (Boyd and
Lagoudas, 1996). The superelasticity model by Auricchio and Taylor
(1997) is employed for the generation of UMAT/Nitinol subroutine repeated for three times with a new specimen used for each test. The
available in Abaqus FEA. In recent years, a few studies have utilized this average value measured from the stress-strain curves at 26 °C was used
built-in subroutine and successfully predicted the true behavior of for the determination of material properties.
shape memory alloy under its operating conditions (Pelton et al., 2013; The modified three point bending test was conducted using the
Robertson et al., 2015; Weafer et al., 2016). same universal testing machine with a customized compressive loading
In this study, a three dimensional numerical modeling of wire- jig, as shown in Fig. 1. The lower load-cell capacity of 500 N was in-
bracket system was developed to anticipate the bending behavior of stalled for this test to increase the sensitivity of load reading. The
superelastic NiTi arch wire. The model adopted the concept introduced clinical scale of three brackets apparatus was set up by setting the
by Auricchio and Taylor (1997) and a nonlinear finite element proce- distance between the brackets to be 7.5 mm, thus resembling the
dure was incorporated into it. The objective of this work is to evaluate average distance of adjacent teeth (Nucera et al., 2014). Three brackets
the influence of binding friction towards force-deflection characteristics with 0.56 slot height and 2.80 mm slot width were used with respect to
of NiTi arch wire with respect to various magnitudes of deflection and its zero torque and angulation design. The brackets were glued to the
oral temperatures. The evolution of normal stress and martensite frac- movable indenter and mounting base using cyanoacrylate adhesives.
tion in the finite element model is subsequently analyzed to understand No ligation tie was installed during the placement of wire inside the
the utilization of superelastic behavior during bending. This new bracket slot. Next, the specimen was deflected to 4 mm at 26 °C by
knowledge may assist orthodontist to quantify the true mechanical re- setting the indenter to move vertically downward and upward at a
sponse of NiTi arch wire in brackets configuration whilst improving displacement rate of 1 mm/min. This bending test was repeated for two
current installation practices. times, and the validation was done by comparing the experimental and
numerical result directly. In this study, the coupling of 0.4 mm ×
2. Materials and methods 0.56 mm wire with the 0.56 mm-slot bracket reflects the common wire-
bracket combination used during the leveling treatment (Lombardo
2.1. Experimental testing et al., 2012). In fact, this 0.16 mm clearance is sufficient to permit
sliding of the arch wire along the bracket slot (Proffit et al., 2014).
In this study, three types of experimental analyses were carried out; The phase transformation temperatures of the specimens were de-
uniaxial tensile, modified three point bending and calorimetric test. The termined using TA-Q20 differential scanning calorimeter. The specimen
uniaxial tensile test was conducted to quantify the mechanical beha- size was 7.86 mg and the heating and cooling rate was 10 °C/min. The
viors parameters of NiTi wire for defining the superelastic behavior in transformation temperatures were determined by the intersection of
the material subroutine. Modified three point bending test was exe- tangent lines on the peaks as described in ISO 15841.
cuted by considering three brackets engagement during the test. The
result of bending test was used to validate the force-deflection curve 2.2. Constitutive model of shape memory alloy
generated from the numerical model. The calorimetric test was done to
measure the transformation temperatures of the arch wire, in specific To simulate the superelasticity of the NiTi arch wire, a built-in user
the austenite finish (Af) temperature. The specimen used for all analyses material subroutine (UMAT/Nitinol) in Abaqus 6.12.2 was employed.
was superelastic NiTi arch wire with a rectangular cross section of This subroutine was established by Auricchio and Taylor (1997), based
0.4 mm × 0.56 mm. on a generalized plasticity theory that decomposes strain components
20 mm uniaxial specimens were cut from the straight end section of into purely elastic and transformational strains. The change in linear
the arch wire. The specimens were elongated to 1.7 mm at a displace- elasticity from the parent austenite phase to the stress-induced mar-
ment rate of 1 mm/min in the direction of its length before unloaded at tensite phase was achieved by adopting a rule of mixtures. The change
similar rate. This test was conducted on an Instron model 3367 uni- in the stress level to induce phase transformation shifts linearly with
versal testing machine. The uniaxial test was carried out at three dif- temperature. This model was chosen over other viable macro scales
ferent temperatures of 26 °C, 36 °C and 46 °C. A small chamber was constitutive models for its good agreement with experimental result
used to control the environmental temperature of the specimen during involving bending type deformation (Nematzadeh and Sadrnezhaad,
tensile test, with an accuracy of ± 1 °C. The desired temperature was 2012).
achieved by continuous flow of heated air through inlet and outlet The material data used in the subroutine are tabulated in Table 1.
vents. The testing procedure and settings followed the standard of ISO These data are generated from the uniaxial test of equiatomic NiTi,
15841: Dentistry-Wires for use in Orthodontic. This uniaxial test was deformed at room temperature.

235
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Table 1 treatments (Parvizi and Rock, 2003). For comparison, the forces exerted
User material data for constitutive model of superelastic NiTi arch wire (T = 26 °C). by the bent arch wire to the central bracket were collected from the
assigned reference point.
Parameter Description Value (unit)

EA Austenite elasticity 44 (GPa) 3. Results


(νA) Austenite Poisson's ratio 0.33
EM Martensite elasticity 23 (GPa)
3.1. Thermal and mechanical deformation behavior
(νM) Martensite Poisson's ratio 0.33
(εL) Transformation strain 0.055
(δσ/δT)L Loading 6.7 (MPa/°C) Fig. 3 displays the thermal martensite phase transformation beha-
σSL Start of transformation loading 377 (MPa) vior of the arch wire specimen. The transformation on cooling occurred
σEL End of transformation loading 430 (MPa) in two stages preceded by the R-phase, thus shifted the martensite
T0 Reference temperature 26 (°C)
transformation temperature to further lower. The reverse transforma-
(δσ/δT)U Unloading 6.7 (MPa/°C)
σSU Start of transformation unloading 200 (MPa) tion on heating occurred in a single stage of martensite to austenite
σEL End of transformation loading 140 (MPa) with heat enthalpy of, Δh = 14.72 J/g. The transformation tempera-
σSCL Start of transformation stress in compression 452 (MPa) tures indicating the start and finish of the martensite phase transfor-
mation were measured at the intersection of tangent lines, as indicated
in the figure. For this specimen the austenite finish, Af is 16.13 °C and
2.3. Finite element simulation
martensite start, Ms is − 50.22 °C. This implies that at room tempera-
ture (~ 26 °C), the structure is at austenite state, thus tensile and
In this study, the generated outputs from the numerical model were
bending deformations at this temperature occur under superelastic
compared against the experimental results. The comparisons were
behavior.
executed to validate the numerical model capability in predicting the
Fig. 4 compares the stress-strain curve generated from the numerical
deformation behavior of superelastic NiTi arch wire under tensile and
against the experimental uniaxial test conducted at three different
bending loads.
temperatures. For clarify purpose, only the most identical curve was
plotted in each plot for direct comparison with the numerical result.
2.3.1. Tensile test The experimental data is indicated by the solid line and the numerical
The 20 mm length uniaxial specimen was modeled by using a total data is indicated by the dashed line. The symbols labelled on Fig. 4a
of 7884 linear hexahedral elements (C3D8). The load was applied at characterized the material data used in Table 1. For all three test
one end by employing a displacement boundary condition while the temperatures, the numerical agreed well with the experimental. The
other end was set to be fixed. The testing parameters and conditions specimen exhibited good superelastic behavior as expressed by the
were replicated from Section 2.1 to preserve the testing integrity. constant stress plateaus of about 6% strain, on both loading and un-
Throughout the test, the total stress reacted on the direction of loading loading. This good agreement between numerical and experimental
was recorded at the fixed end of the specimen. results, in terms of loading and unloading plateau, transformation,
elastic and residual strains provided a good confidence for extended
2.3.2. Bending test modeling of deformation behavior of the alloy at different boundaries
The 30 mm length bending specimen was modeled by using a total and parameters conditions. Fig. 4d summarizes the temperature de-
of 71,144 linear hexahedral elements with reduced integration pendent of stress-induced martensite transformation against tempera-
(C3D8R). The hour glassing effect of this reduced integration second ture. The slope of stress increase against temperature was recorded to
order eight-node element was controlled by the selection of stiffness be 6.7 MPa/°C, agreed well with the Classius-Clayperon relation
hourglass behavior (ABAQUS manual). The arch wire specimen was (Miyazaki and Otsuka, 1989). It was also acknowledged that a small
assumed to be mechanically isotropic. A convergence analysis of force- variation was registered in the value of critical stress for transforma-
deflection response was performed to select an appropriate elements tion, as indicated by the error bars inserted in the curves.
size for the model. The bending model was considered to be converged Fig. 5 displays the force-deflection plot of NiTi specimen upon
when the deviation of force between the numerical and experimental 4.0 mm bending on three brackets system at 26 °C. The activation cycle
results was less than 10%. began with a short linear slope up to 0.3 mm of deflection, representing
The bracket was designed into two parts, bottom and top part by the bending stiffness of the arch wire just before making a contact with
using bilinear rigid quadrilateral element (R3D4). Both parts have a the top lateral edges of the bracket. Once this contact was established,
width of 2.80 mm and were separated by a distance of 0.56 mm, re- the wire became more constrained within the bracket slot, thus yielded
presenting the real bracket slot dimension. The arch wire was as- a steeper slope before reaching the plateau at about 1.1 mm of deflec-
sembled into the brackets slot with no boundary condition defined to let tion length. The deactivation cycle started with a sudden force drop,
the wire to slide freely during the bending. Contact interaction between implying the onset of martensite transformation with regards to the
arch wire and bracket was set as surface-to-surface contact with a stress relief. Gradual force increment was observed throughout the wire
friction coefficient of 0.27. It has been reported in the literature that the recovery from 3 mm to 1 mm. This is believed due to the weakening of
coefficient of friction between NiTi arch wire and stainless steel bracket binding intensity as wire deflection reduces. The force-deflection curve
surface to be in the range of 0.23–0.30 (Whitley and Kusy, 2007). from the simulation, as indicated by the dashed line is in good agree-
As shown in Fig. 2, each bracket was assigned to its own reference ment with the experimental data. This provides a high level of con-
point (RP). Only the central bracket was set to move in y-direction at a fidence on the numerical bending model for prediction of the force
displacement rate of 1 mm/min, while the adjacent brackets were performance of superelastic NiTi arch wire bent in a brackets config-
constrained to be fixed in all directions (Ux = Uy = Uz = 0). The wire uration. Additionally, good consistency was established between the
was set to be deflected to 1, 2, 3 and 4 mm at three different tem- experimental and the numerical curves, as indicated by the small dis-
peratures of 26 °C, 36 °C and 46 °C, respectively. The three tempera- crepancy of force magnitude (0.2 N) at 2.5 mm deflection.
tures used in the present study portray the deformation behavior of the The predicted force-deflection curve of NiTi specimen bent at dif-
wire for different cases; behavior at the time of installation (26 °C), ferent deflection magnitudes, and temperatures are shown in Fig. 6. The
behavior inside the oral environment (36 °C) and behavior during ex- highest magnitudes of activation and deactivation forces were recorded
posure to warm food intake (46 °C). The deflection ranges were chosen at 4 mm and 1 mm, respectively. For a small deflection such as 1 mm,
based on norm clinical conditions encountered during the orthodontic the arch wire was deformed and recovered elastically. A short force

236
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Fig. 2. Finite element model for the modified three point bending test with finer mesh near the contact interaction.

Fig. 3. Thermal transformation behavior of NiTi arch wire.


Fig. 5. Force-deflection curve of NiTi arch wire in a modified three point bending setup.

Fig. 4. Stress-strain curves of NiTi arch wire under uniaxial test at: (a) 26 °C, (b) 36 °C, (c) 46 °C and (d) critical stress for stress-induced martensitic phase transformation.

237
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Fig. 6. Numerical result of force-deflection behavior of NiTi arch wire bent at different deflection magnitudes and temperatures of: (a) 26 °C, (b) 36 °C, (c) 46 °C and (d) comparison of
activation and deactivation plateaus for 4 mm deflection group.

stress needed for martensite to austenite transformation to a higher


extent, thus increasing the gap with the deactivation force. The large
gap between these two forces caused the deactivation rate to increase to
assist the wire recovery process.
It is interesting to note that the deactivation force varies with re-
spect to the maximum deflection applied on the arch wire. As shown in
Fig. 6a, the wire deflected to greater magnitude of deflection recorded
lower deactivation force than the one deflected to smaller magnitude.
This discrepancy in force magnitude formed a force gap (ΔF), which
Fig. 7. Force gap of NiTi arch wire measured at different deflection magnitudes and continues for at least 0.5 mm before the two deactivation curves re-
temperatures. merged. In current practice, orthodontist takes advantage of this force
gap in a reactivation process, by which arch wire is detached from a
bracket and subsequently reattached to improve the deactivation
plateau of austenite to martensite transformation is observable only
magnitude.
during the recovery of arch wire from the 2 mm activation. Beyond
Fig. 7 summarizes the force gap (ΔF) magnitude recorded at dif-
2 mm, the activation and deactivation of the wire lead to a positive and
ferent set of deflections and temperatures. Each deflection group shows
negative gradient plateau, respectively. At a higher temperature, both
a reduction in the magnitude of force gap as the temperature increases
gradient plateaus delayed to a higher magnitude.
from 26 °C to 46 °C. The highest reduction is observed on 1 mm group,
Fig. 6d compares the slope of the force plateaus upon 4 mm de-
with the total difference in the force gap is about 0.73 N. Meanwhile,
flection at different temperatures. Based on the linearity of the gradient
the lowest reduction is logged on 3 mm group, with only 0.07 N of force
plateaus in Figs. 6a, 6b and 6c, the slope is selected to be from 2 to
magnitude difference. This trend shows that the force gap reduces at
4 mm during the activation and from 3 to 1 mm during the deactiva-
higher deflection group, and the change of temperature starts to be less
tion, respectively. It is observed that the arch wire shared similar ac-
sensitive to the force gap magnitude.
tivation rate of 1.0 N/mm along the selected deflection interval, re-
gardless of the applied wire temperature. This similar activation rate
corresponds to the identical increase of stress needed for martensitic 3.2. Finite element results
transformation and binding magnitude against the 10 °C temperature
elevation. In contrast, the deactivation occurred at dissimilar rates, Fig. 8 displays the normal stress distribution (S11) of the arch wire
where the steepest force plateau of 1.1 N/mm occurred at 46 °C, fol- during bending activation in 36 °C environment. Red and blue contour
lowed by 0.88 N/mm and 0.66 N/mm at 36 °C and 26 °C, respectively. expressed the elements under tension and compression, respectively.
At 46 °C, the thermomechanical behavior of the wire increased the The peak of compressive stress is much higher than the tensile stress
due to the greater compressive critical stress set for the model. The

238
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Fig. 8. Normal stress distribution (S11) of NiTi arch


wire upon 4 mm activation at 36 °C.

contour located at the bending curvature near to the bracket corner located at the core and the remaining region of the wire.
characterizes the localized martensitic transformation. The maximum
value of tensile normal stress is observed on three hotspot regions,
specified to the outer elements of the wire that experienced the most 4. Discussions
bending deformation.
Fig. 9 summarizes the maximum S11 magnitude measured during 4.1. Bending behavior in bracket configuration
wire bending in 36 °C environment. Along the 4 mm activation, the
maximum S11 was elevated from 455 to 562 MPa. It is interesting to In this study, the bending behavior of NiTi arch wire at different
note that at 3 mm, the maximum S11 (551 MPa) exceeded the end vertical deflections was simulated by three dimensional numerical
stress level of stress-induced martensitic transformation (σEL), which is modeling. The binding generated at the arch wire-bracket contact was
approximately 480 MPa (obtained from Fig. 6b). This stress discrepancy considered to capture the influence of this sliding resistance component
accentuated that further wire activation to above 2 mm was accom- towards bending behavior. The bending behavior was evaluated by
panied by the elastic straining of martensite structure. Notice that the observing the force-deflection trend, martensite fraction and maximum
deactivation from 4 to 2 mm corresponds to the combination of elastic normal stress developed during vertical bracket displacement. The
and martensitic transformation. combination of experimental and numerical visualization provides a
Fig. 10 shows the volume fraction of martensite structure (SDV21) clear insight of how superelastic behavior of NiTi arch wires is deployed
upon 4 mm activation at 36 °C. SDV21 is a nodal output of the sub- during the orthodontic treatment.
routine, represents the volume fraction of austenite transformed to Referring to Fig. 4, the arch wire elongated at temperature above Af
martensite structure. The 0.0 and 1.0 magnitude denotes the zero and exhibited stress-strain curve in the form of superelastic behavior. The
fully transformed martensite region with respect to the bending de- linear elastic deformation at ~ 1% strain reflects the elastic yield of the
formation of the wire. It is seen that NiTi wire evolves in three stages: a austenite structure. As the stress is further increased, the deformation
fully transformed martensite (M) zone located at the outer top portion exhibited a stress plateau with positive gradient. At this stage, the wire
of the bent region, a partial transformation zone (A–M) that surrounded deformation involved microscopic phase change from austenite struc-
the fully transformed martensite zone, and a single austenite (A) phase ture into martensite via SIM transformation. Surpassing this plateau,
the elongation proceeds with linear stress elevation, imitates the elastic
yielding of martensite structure. Upon unloading, this stress-induced
martensite variant becomes unstable and begins to reverse back to its
original phase structure, austenite at a lower stress level, thus forming a
stress hysteresis.
Similar deformation trend is expected to happen during bending.
Unfortunately, the activation and deactivation force measured from the
modified three point bending does not exhibit a similar plateau as de-
formed in three point bending (Wilkinson et al., 2002) and uniaxial test.
Instead, the plateau region exhibited positive and negative gradient
slope on the activation and deactivation of the wire, respectively. The
gradient plateau trend has been reported before by Naceur et al. (2014)
who predicted numerically the force-deflection behavior of NiTi arch
wire via bent deformation. They claimed that the gradient force pla-
teaus were formed due to the rise of stress needed for martensitic
transformation with respect to the accumulation of elastic strain en-
ergy.
Fig. 9. Maximum normal stress (S11) measured upon activation and deactivation of NiTi However, in this present study, we believed that these gradient
arch wire at 36 °C.
force plateaus were triggered by the variation of binding friction

239
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

Fig. 10. View cut of martensitic volume fraction (SDV21) evolution in arch wire upon 4 mm activation at 36 °C.

generated at the adjacent brackets such inferred by Baccetti et al. structure. Knowing that superelastic material to exert constant force,
(2009) and Nucera et al. (2014). Both authors did not provide the the regular increase of binding from 1.1 mm and onwards signifies the
binding–deflection data from the adjacent brackets to support their elevation of normal force at the contacting surfaces. This linear trend of
theory due to the limitation of the experimental capabilities. It is sug- binding elevation over deflection agrees well with the experimental
gested that the activation force needed to bend and slide the arch wire work conducted by Murayama et al. (2013) and Reznikov et al. (2010).
increases with respect to the rise of binding friction generated at the
wire-bracket contact. At the onset of wire recovery, the wire spring
4.2. Clinical perspectives
back is used up to overcome the binding friction built up at the contact
regions, thus delaying the deactivation force. As the wire deflection
Even though bending of NiTi arch wire involves activation and
reduces further, this force is gradually elevated signifying the weak-
deactivation cycle, it is only the force measured during the deactivation
ening of the binding.
quantifies the magnitude of force released to the bracket (Elayyan et al.,
The variation of binding intensity over wire deflection at 36 °C is
2010). It is important to understand the force–deflection behavior of
shown in Fig. 11. This binding-deflection plot presents the sliding re-
NiTi arch wire in bracket configuration so that orthodontist can cor-
sistance encountered by the wire following its continuous contact with
relate the magnitude of deactivation force with current wire deflection.
the bottom region of the bracket surfaces. The binding intensity was
The consensus of NiTi material delivering a constant force may no
obtained by multiplying the total magnitude of contact force (CFTM)
longer be true for the case of bracket bending, due to the fact that the
with the defined coefficient of friction. The linear increase of binding
wire delivers force in a gradient manner.
from 0 to 1.1 mm was attributed to the elastic yielding of the austenite
In the present study, the discussion only reflects the bending of
0.4 mm × 0.56 mm NiTi wire with 7.5 mm inter bracket setting. The
preliminary data here provides an early insight on how changes in wire
stiffness (owing to temperature changes) influenced the force magni-
tude of the wire. Besides the changes in the oral temperature, the or-
thodontic wire may also be subjected to bend with a smaller inter
bracket configuration. From an extended simulation work, if the same
wire is to be bent at 1.0 mm with smaller inter bracket distance, per say
7.0 mm, the activation and deactivation force was further increased
approximately by 23%. These force increments were highly related
with the increased in wire stiffness corresponded to the shortening of
wire length for active bending.
The trend of arch wire to release different deactivation force with
respect to the maximum deflection applied could be clinically relevant
for reactivation process. This reactivation process has been used by
orthodontist to elevate the amount of force released by the superelastic
NiTi wire that had previously been activated. For example; in the case
of tooth that has moved by 2 mm from its original 4 mm bracket dis-
placement, reactivating the same wire at 2 mm displacement may
Fig. 11. Prediction of binding friction of wire bent at 36 °C.
trigger higher deactivation force than the one transmitted before. A

240
M.F. Razali et al. Journal of the Mechanical Behavior of Biomedical Materials 77 (2018) 234–241

data on how much this reactivation process improves the deactivation NiTi wire occurred locally, concentrated at the region of highly
force is necessary so that this process can be executed with precaution. deformed structure.
Since binding influences the deactivation forces, more study on the
force gap for different wire-bracket combination at various bending Acknowledgments
configuration is a necessity.
During the orthodontic treatment period, it is a common practice for The authors are grateful for the financial support provided by
orthodontist to gradually change the geometry of arch wire from round Universiti Sains Malaysia under the grant RUI 1001/PMEKANIK/
to rectangular shape. This procedure is practiced to fill up the slot of the 814244.
bracket so that the teeth can move into predetermined position with
respect to the tip and torque prescribed on the bracket geometry. It is References
interesting to highlight that for every loading condition executed in this
study, the combination of 0.4 mm × 0.56 mm rectangular wire with a Auricchio, F., Taylor, R.L., 1997. Shape-memory alloys: modelling and numerical simu-
lations of the finite-strain superelastic behavior. Comput. Methods Appl. Mech. Eng.
0.56 mm slot height bracket is ineffective to provide suitable deacti-
143, 175–194.
vation force (< 1.0 N) for tooth movement. Referring to Fig. 6b, the Baccetti, T., Franchi, L., Camporesi, M., Defraia, E., Barbato, E., 2009. Forces produced by
rectangular wire may generate as much as 5.0 N force for a 1 mm different nonconventional bracket or ligature systems during alignment of apically
displaced teeth. Angle Orthod. 79 (3), 533–539.
bracket displacement. This deactivation force is too high for human Boyd, J.G., Lagoudas, D.C., 1996. A thermodynamical constitutive model for shape
tissue and may cause hyalinization of periodontal ligament, hence memory materials. Part I. The monolithic shape memory alloy. Int. J. Plast. 12,
discourages effective tooth movement (Proffit et al., 2014). Therefore, 805–842.
Duerig, T.W., Melton, K.N., Stöckel, D., 2013. Engineering Aspects of Shape Memory
the current practice of using thick rectangular wire at the end of le- Alloys. Butterworth-Heinemann, London.
veling treatment should be revised to comply with the ideal force cri- Elayyan, F., Silikas, N., Bearn, D., 2010. Mechanical properties of coated superelastic
archwires in conventional and self-ligating orthodontic brackets. Am. J. Orthod.
teria. Dentofac. Orthop. 137, 213–217.
Apart from the light force criteria, effective tooth movement also Jiang, F., Liu, Y., Yang, H., Li, L., Zheng, Y., 2009. Effect of ageing treatment on the
requires the force to be delivered in a constant magnitude throughout deformation behaviour of Ti–50.9 at% Ni. Acta Mater. 57, 4773–4781.
Kaphoor, A.A., Sundareswaran, S., 2012. Aesthetic nickel titanium wires–how much do
the treatment period. On account of faster tooth movement over a they deliver? Eur. J. Orthod. 34, 603–609.
continuous application of force (Weiland, 2003), minimizing the slope Kim, T.-K., Kim, K.-D., Baek, S.-H., 2008. Comparison of frictional forces during the initial
gradient is necessary to achieve shorter treatment duration. For this leveling stage in various combinations of self-ligating brackets and archwires with a
custom-designed typodont system. Am. J. Orthod. Dentofac. Orthop. 133 (187-e15).
reason, reduction on friction coefficient and normal force with which Kusy, R., Whitley, J., 1999. Influence of archwire and bracket dimensions on sliding
the contacting surfaces are pressed together is compulsory to reduce the mechanics: derivations and determinations of the critical contact angles for binding.
Eur. J. Orthod. 21, 199–208.
sliding resistance, and hence its impact on the force-deflection trend. Lombardo, L., Marafioti, M., Stefanoni, F., Mollica, F., Siciliani, G., 2012. Load deflection
Thus, due to the direct relation between wire stiffness and binding characteristics and force level of nickel titanium initial archwires. Angle Orthod. 82,
(Tageldin et al., 2016), selecting a smaller wire might be helpful to 507–521.
Miyazaki, S., Otsuka, K., 1989. Development of shape memory alloys. ISIJ Int. 29,
minimize the magnitude of normal force and to lessen the binding in- 353–377.
tensity. Meanwhile, applying a special surface treatment on bracket and Murayama, M., Namura, Y., Tamura, T., Iwai, H., Shimizu, N., 2013. Relationship be-
tween friction force and orthodontic force at the leveling stage using a coated wire. J.
arch wire surfaces may be a pleasant preference in reducing the friction
Appl. Oral Sci. 21 (6), 554–559.
coefficient. Both options may be a decent area to be explored in the Naceur, I. Ben, Charfi, A., Bouraoui, T., Elleuch, K., 2014. Finite element modeling of
future, to preserve light and constant force behavior of superelastic NiTi superelastic nickel-titanium orthodontic wires. J. Biomech. 47, 3630–3638.
Nematzadeh, F., Sadrnezhaad, S.K., 2012. Effects of material properties on mechanical
arch wire in the bracket system. performance of Nitinol stent designed for femoral artery: finite element analysis. Sci.
The deactivation force of NiTi wire is highly dependent on tem- Iran. 19, 1564–1571.
perature. The force increases as the oral temperature increases. This Nucera, R., Gatto, E., Borsellino, C., Aceto, P., Fabiano, F., Matarese, G., Perillo, L.,
Cordasco, G., 2014. Influence of bracket-slot design on the forces released by su-
implies that intake of warm water by a patient leads to the stiffening of perelastic nickel-titanium alignment wires in different deflection configurations.
NiTi wire, thus greater force is expected to be delivered to the mal- Angle Orthod. 84, 541–547.
Otsuka, K., Wayman, C.M., 1999. Shape Memory Materials. Cambridge University Press,
occlusion. For this reason, if the deactivation force of a specific wire at Cambridge.
body temperature is estimated to be close to the 1.0 N, intake of warm Parvizi, F., Rock, W.P., 2003. The load/deflection characteristics of thermally activated
drinks should be avoided to evade pain. Additionally, NiTi wire also orthodontic archwires. Eur. J. Orthod. 25, 417–421.
Pelton, A.R., Fino-Decker, J., Vien, L., Bonsignore, C., Saffari, P., Launey, M., Mitchell,
shows greater variation in the magnitude of deactivation force in high M.R., 2013. Rotary-bending fatigue characteristics of medical-grade Nitinol wire. J.
temperature environment. Mech. Behav. Biomed. Mater. 27, 19–32.
Proffit, W.R., Fields Jr, H.W., Sarver, D.M., 2014. Contemporary Orthodontics, fifth ed.
Mosby Press, St. Louis.
5. Conclusions Reznikov, N., Har-Zion, G., Barkana, I., Abed, Y., Redlich, M., 2010. Measurement of
friction forces between stainless steel wires and “reduced-friction” self-ligating
brackets. Am. J. Orthod. Dentofac. Orthop. 138, 330–338.
The present study investigates the bending behavior of rectangular Robertson, S.W., Launey, M., Shelley, O., Ong, I., Vien, L., Senthilnathan, K., Saffari, P.,
NiTi arch wire at different magnitude of deflections and temperatures. Schlegel, S., Pelton, A.R., 2015. A statistical approach to understand the role of in-
The following conclusions can be deduced from this study: clusions on the fatigue resistance of superelastic Nitinol wire and tubing. J. Mech.
Behav. Biomed. Mater. 51, 119–131.
Singh, G., 2015. Textbook of Orthodontics, third ed. Jaypee Brothers Medical Publishers,
1. The wire subjected to greater deflection magnitude released lower New Dehli.
Tageldin, H., de Llano Perula, M., Thevissen, P., Celis, J.-P., Willems, G., 2016. Resistance
magnitude of deactivation force as compared to the one deflected at
to sliding in orthodontics: a systematic review. Jacobs J. Dent. Res. 3, 34.
smaller deflection, thus creating a force gap. The force gap value Varela, J.C., Velo, M., Espinar, E., Llamas, J.M., Rúperez, E., Manero, J.M., Javier Gil, F.,
reduced as the wire temperature and activation increased. 2014. Mechanical properties of a new thermoplastic polymer orthodontic archwire.
Mater. Sci. Eng. C Mater. Biol. Appl. 42, 1–6.
2. The variation of binding intensity at bracket contacts influenced the Weafer, F.M., Guo, Y., Bruzzi, M.S., 2016. The effect of crystallographic texture on stress-
force-deflection characteristics of NiTi wires by elevating and de- induced martensitic transformation in NiTi: a computational analysis. J. Mech.
laying the activation and deactivation force, respectively. Behav. Biomed. Mater. 53, 210–217.
Weiland, F., 2003. Constant versus dissipating forces in orthodontics: the effect on initial
3. The wire exhibited greater activation and deactivation force at tooth movement and root resorption. Eur. J. Orthod. 25, 335–342.
higher bending temperature owing to the enhancement of wire Whitley, J.Q., Kusy, R.P., 2007. Influence of interbracket distances on the resistance to
sliding of orthodontic appliances. Am. J. Orthod. Dentofac. Orthop. 132, 360–372.
thermomechanical properties. Wilkinson, P.D., Dysart, P.S., Hood, J.A.A., Herbison, G.P., 2002. Load-deflection char-
4. The finite element analysis demonstrated that the activation of NiTi acteristics of superelastic nickel-titanium orthodontic wires. Am. J. Orthod. Dentofac.
wires beyond 2 mm in bracket configuration was accompanied by Orthop. 121, 483–495.
martensitic elastic deformation. The martensitic transformation of

241

You might also like