Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

ORIGINAL ARTICLE

Load-deflection characteristics of superelastic


nickel-titanium orthodontic wires
Peter D. Wilkinson, MDS,a Peter S. Dysart, MDS,b James A. A. Hood, ED BSc, MDS, FRACDS,c and
G. Peter Herbison, MScd
Dunedin, New Zealand

Previous mechanical testing of orthodontic wires has, in many cases, failed to simulate some key features of
the clinical environment. The purpose of this study was to investigate the load-deflection characteristics of
7 different 0.016-in initial alignment archwires (Twistflex, NiTi, and 5 brands of heat-activated superelastic
nickel-titanium [HASN]) with modified bending tests simulating a number of conditions encountered clinically.
Load-deflection tests were carried out on the wires with 5 different model designs, and data from selected
points on the unloading phase of the generated graphs were statistically analyzed. Wire deflection was
carried out at 3 temperatures (22.0°C, 35.5°C, and 44.0°C) and to 4 deflection distances (1 mm, 2 mm, 3 mm,
and 4 mm). Rankings were derived according to statistically significant differences in each test situation. The
effects of model, wire, and temperature variation were all statistically significant. Twistflex and the 5 HASN
wires produced a range of broadly comparable results, and NiTi gave the highest unloading values. Model
rankings indicated that self-ligating Twin-Lock brackets produced lower friction than regular edgewise
brackets. The authors recommend using the rankings from the mechanical test simulations to predict
possible clinical performance of archwires. (Am J Orthod Dentofacial Orthop 2002;121:483-95)

I
n the last 10 to 15 years, superelastic nickel- from the body centered cubic austenitic form to the
titanium archwires have gained acceptance by monoclinic martensitic form of nickel-titanium when
orthodontists as initial alignment archwires, largely the stress reaches a certain level during activation and
because of their unique properties of superelasticity and deactivation.1,7,12 In addition to the austenitic and
shape memory.1-3 Most of the information about the martensitic phases, an intermediate phase with rhom-
behavior of these wires is based on mechanical labora- boidal symmetry has been described as the R phase.12
tory testing with an emphasis on 3-point bending tests Alloys that can be plastically deformed in their
to study load-deflection characteristics4-8 without sim- martensitic phase, but will return to an austenitic phase
ulating the many variables encountered in clinical and thereby recover their original shapes if heated
situations.3 through a certain transition temperature range (TTR),
Superelasticity is characterized by a load-deflection are said to exhibit shape memory. Early superelastic
plot with a horizontal region during unloading, imply- nickel-titanium wires exhibited shape memory charac-
ing that a constant force may be exerted over that teristics, but their TTRs made it impractical to exploit
particular range of tooth movement.1,9,10 Some authors this property for orthodontic treatment.13 More re-
prefer to describe such behavior as pseudoelastic- cently, third and fourth generation temperature-depen-
ity.11,12 Metallurgical studies have attributed these dent heat activated (also termed thermoresponsive or
characteristics to a reversible phase transformation thermodynamic) nickel-titanium wires have been mar-
keted with clinically useful shape memory.
From the University of Otago, Dunedin, New Zealand.
a
Graduate, Discipline of Orthodontics, Department of Oral Sciences and Three-point bending tests offer reproducibility,
Orthodontics. which has facilitated comparison between studies.5,6,14
b
Senior Clinical Lecturer, Discipline of Orthodontics, Department of Oral However, the appropriateness of such a simple test is
Sciences and Orthodontics, and in private practice.
c
Professor, Discipline of Biomaterials, Department of Oral Sciences and called into question by the failure of clinical trials to
Orthodontics. confirm that superelastic nickel-titanium wires outper-
d
Senior Research Fellow, Department of Preventive and Social Medicine. form less expensive nitinol or multistrand stainless steel
Reprint requests to: Professor J. A. A. Hood, Department of Oral Sciences and
Orthodontics, PO Box 647, Dunedin, New Zealand; e-mail, jim.hood wires during initial alignment.15-17 It has been sug-
@dent.otago.ac.nz. gested that the most appropriate wire tests may there-
Submitted, January 2001; revised and accepted, July 2001. fore be those that reproduce conditions encountered
Copyright © 2002 by the American Association of Orthodontists.
0889-5406/2002/$35.00 ⫹ 0 8/1/121819 clinically, with the wire constrained as part of a fixed
doi:10.1067/mod.2002.121819 appliance.18 Variations in model design have been
483
484 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

Table I. Orthodontic 0.016-in round wires tested

Wire Supplier Location

Twistflex GAC International Islandia, NY


NiTi Ormco Glendora, Calif
Copper NiTi 27 Ormco Glendora, Calif
(27 CuNiTi)
Copper NiTi 35 Ormco Glendora, Calif
(35 CuNiTi) Fig 1. Interbracket distance (mm) used on models
Reflex TP Orthodontics Inc. La Porte, Ind (modified after Moyers et al28).
Tensic Dentaurum Ispringen, Germany
Nitinol HA 3M Unitek Monrovia, Calif

Table II. Models

Model Description

1 Three-point bending with brass jig


2 Partial acrylic block model with 4 Mini-Diamond
brackets
3 Partial acrylic block model with 4 Twin-Lock
self-ligating brackets
4 Full acrylic arch form with maxillary arch Mini-Diamond
brackets
5 Full acrylic arch form with maxillary arch Twin-Lock
brackets

Fig 2. Three-point bending (model 1).


2,19
shown to affect unloading deflection plots, and this
must be considered when analyzing data. Other factors accordance with these average tooth widths (Fig 1) to
that may alter the performance of wires include tem- simulate either a full or a partial arch. Brackets were
perature5,20-22 and friction between the archwire and placed with their long axes parallel in all acrylic models
the bracket.23-27 and in the same positions on the 2 arch-form models.
The purpose of this study was to investigate the
load-deflection characteristics of superelastic nickel- Model designs
titanium and nonsuperelastic initial alignment wires For the 3-point bend design (Fig 2), the wire sample
with modified bending tests. was held in a brass jig with supports 15.5 mm apart and
displaced at the midpoint.
MATERIAL AND METHODS For model designs 2 and 3, partial acrylic block
A selection of 0.016-in round orthodontic wires was (Fig 3), 2 rows (1 of Mini-Diamond brackets and 1 of
tested ex vivo to compare their mechanical properties. Twin-Lock brackets [both, Ormco, Glendora, Calif]) of
The sample comprised 7 brands: 5 were marketed as 4 brackets were cemented to the partial acrylic block.
heat activated superelastic nickel-titanium (HASN) Brackets were placed in positions representing teeth 11,
and 1 as superelastic nickel-titanium, and the remaining 21, 23, and 24 with the load cell acting as the displaced
wire was multistrand stainless steel (Table I). The wires lateral incisor. For the first 2 model designs, 50.0 mm
were subjected to 3-point bending tests and were also sections of straight wire were tested.
tested with 2 types of orthodontic brackets on 2 designs For designs 4 and 5, full acrylic arch form (Fig 4),
of acrylic models, giving a total of 5 test models (Table 2 arch forms were constructed, both matching the
II). In each situation, the load site simulated a mal- Orthos Small Maxillary Archform template (Ormco).
aligned lateral incisor with 15.5 mm between the To 1 model, a set of Mini-Diamond maxillary brackets
supports of the brass jig, or between the midpoints of was cemented with adhesive (GAC International, Is-
the brackets. This interbracket distance was derived landia, NY), while Twin-Lock brackets were placed on
from typical tooth dimensions (accurate to within 0.5 the other. Specifications of the brackets are given in
mm) for a male’s maxillary permanent dentition.28 On Table III. Each model had identical molar tubes
the acrylic models, the brackets were positioned in (Ormco) attached at the sites of teeth 16 and 26, and the
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 485
Volume 121, Number 5

slot size for all brackets and tubes was 0.022 ⫻ 0.028
in. All preformed archwires on these models were
selected to resemble Orthos Small Maxillary Archform.
In the models with Mini-Diamond brackets,
0.110-in elastomeric PowerOModules (Ormco) liga-
tures were placed with a Straight Shooter ligature gun
(TP Orthodontics, LaPorte, Ind) to minimize variation
in the degree of stretching during application. The
Twin-Lock brackets were self-ligating. Each model
was then placed in the same position in a bending jig
with locating devices.

Bending method
A modified bending test was carried out on the
wires with an Instron 1193 universal testing machine
(Instron Corp, Canton, Mass) fitted with a 100-kg
compression load cell calibrated on the 2-kg range with Fig 3. Partial acrylic block (models 2 and 3).
a 2-kg standard weight traceable to the National Bureau
of Standards of the United States and a deflecting rod
tip formed from 1.6-mm diameter polished steel (Fig
5). The testing machine was operated at a crosshead
speed of 2.0 mm per minute, and the analogue output
signal was passed through a MacLab/4 analogue-digital
converter (A. D. Instruments, Castle Hill, Australia) to
a computer (Macintosh LCII, Apple, Cupertino, Calif).
Displacement was measured by a model GCA-121-250
LVDT (Lucas Schaevitz, Hampton, Va) calibrated
against a differential micrometer model 110-120 (Mi-
tutoyo Corp, Tokyo, Japan), reading to 0.0001 mm.
Each bending test was carried out 3 times, with a new
piece of wire for each repetition. The wire was loaded
in a buccolingual direction in the flat acrylic block
model and gingivo-occlusally in the arch-form model to
allow greater stability and positional accuracy. It has Fig 4. Full acrylic arch form (models 4 and 5).
been shown in previous studies that such differences in
loading direction do not alter mechanical behavior.19,29
Wires were tested at temperatures of 22.0°C, 35.5°C, measured at 1.0 and 0.5 mm. For the remainder, the
and 44.0°C. Wire temperature was controlled by im- deflection points chosen were 0.5 mm less than the
mersing the bending jig containing the model and the maximum deflection of each bend and 1.0 mm (Fig 6).
wire specimen in a brass water bath attached to the
Instron load cell. The water solution was exchanged Statistical analysis
from a large precalibrated bath, and temperature at the The raw data values were transferred to a spread-
bending jig was verified with a precalibrated K-type sheet and, with a software package (SPSS for Windows
thermocouple. Wires were deflected to 1, 2, 3, and 4 98, version 8.0, SPSS Inc, Chicago, Ill), an analysis of
mm in separate load-deflection tests. variance (ANOVA) was then used to separate the
different effects. Post hoc analyses, with the Duncan
Measurements multiple range test, were applied to analyze the main
Load value measurements from the unloading effect of different wires (keeping the different model
phase were derived digitally by selecting 2 deflection designs and temperature constant) and models (keeping
points on an X-Y plot on the computer screen. These the different wire types and temperature constant). P
points were chosen to represent a standardized portion ⬍.05 denoted statistical significance. Rankings were
of the unloading phase plateau in the load-deflection assigned to the wire and the model values based on the
curve. For tests deflected to 1 mm, the load values were results of the statistical analysis.
486 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

Table III. Orthodontic brackets (stainless steel, slot size 0.022 ⫻ 0.028-in) in maxillary arch

Torque Angulation In/out Mes/dist Occ/ging


Bracket Tooth (°) (°) (mm) (mm) (mm)

Mini-Diamond Central ⫹14 ⫹5 0.8 3.75 3.0


Lateral ⫹7 ⫹8 1.0 3.0 3.0
Canine 0 ⫹10 0.5 3.25 3.0
1st premolar ⫺7 0 0.5 3.125 3.0
2nd premolar ⫺7 0 0.5 3.125 3.0
Twin-Lock Central ⫹14 ⫹5 0.7 3.0 3.125
Lateral ⫹7 ⫹8 0.9 3.0 3.125
Canine 0 ⫹10 0.5 3.0 3.125
1st premolar ⫺7 0 0.6 3.0 3.125
2nd premolar ⫺7 0 0.6 3.0 3.125

Mes/dist, Mesiodistal; Occ/ging, occlusal-gingival.

and method of attachment. The unloading plot for


3-point bending tests typically dropped very rapidly
followed by a long range of deactivation during which
a relatively constant force was produced (Fig 7, Graph
1 for 27CuNiTi). In this superelastic range, the load
curves for activation and deactivation were consistent
with the definition of hysteresis.21
The results of the ANOVA show that the effects of
model, wire, and temperature variation were all signif-
icantly different at the P ⬍ .05 level.

Effect of model design


Fig 5. Three-point bending model in temperature-con- Table IV presents results from the load-deflection
trolled water bath in testing machine. curves during unloading in these clinically relevant
conditions. A mean load value is given for each of the
2 points on the curve, and the plateau gap is the
average load difference between the value at the 1.5-
mm unloading deflection point (UDP) and the 1.0-mm
UDP. It therefore represents a measure of gradient of
the unloading plateau over a standardized deflection
distance. Load values varied depending on the model
used; in model 2, the 1.5-mm UDP values ranged from
23.33 to 120.00 g for the 5 HASN wires and, in model
5, from 103.30 to 166.70 g. NiTi (Ormco) consistently
had the highest load value at each UDP in every model.
The 35 CuNiTi (Ormco) wire consistently had low
values and, in model 4, produced zero load at both 1.5
Fig 6. Typical X-Y output for superelastic nickel-tita- and 1.0 mm.
nium wire at 2-mm load-deflection test. UDP, Unloading Table V presents the results of comparing the
deflection point; ELDP, end load deflection point. different wires while keeping the different model de-
signs and temperatures constant. Included is the end
load deflection point (ELDP), representing the mean
RESULTS deflection value at which the load reached zero in the
Most, but not all, load-deflection graphs of the unloading phase. Each wire is ranked from 1 to 7 for
superelastic NiTi wires confirmed features of superelas- each measurement, with 1 the lowest value and 7 the
ticity, with plateau regions varying in extent, gradient, highest, based on statistically significant differences
and load value depending on wire, model, temperature, shown by the ANOVA. When 2 measurements were
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 487
Volume 121, Number 5

not statistically different, they were allocated the same Load values (g) during unloading from 2.0-
Table IV.
rank (⫽). NiTi wire attained the highest ranked load mm deflection at 35.5°C
values at the 1.5-mm UDP and 1.0-mm UDP, while
Load (g) at Load (g) at
Twistflex (GAC International) produced values compa- 1.5-mm UDP 1.0-mm UDP
rable to the HASN wires. The ranking of the 5 HASN Plateau
wires remained similar for both UDPs with Reflex (TP Wire Mean (⫾SD) Mean (⫾SD) gap (g)
Orthodontics) and 27 CuNiTi (Ormco) producing the Model 1
highest ranked load values at the 1.5-mm UDP and 35 Twistflex 36.67 (5.77) 23.33 (5.77) 13.34
CuNiTi the lowest. Twistflex, 35 CuNiTi, and Tensic NiTi 130.00 (0.00) 110.00 (10.00) 20.00
(Dentaurum, Ispringen, Germany) had the highest 27 CuNiTi 73.33 (5.77) 70.00 (0.00) 3.33
35 CuNiTi 43.33 (5.77) 30.00 (0.00) 13.33
ranked ELDPs and delivered zero load below 0.3-mm
Reflex 70.00 (0.00) 50.00 (0.00) 20.00
deflection. Tensic 60.00 (0.00) 46.67 (5.77) 13.33
The effect of model design is summarized in Table Nitinol HA 40.00 (0.00) 20.00 (0.00) 20.00
VI, which shows that load values varied significantly Model 2
between models for some measurements, consistently Twistflex 46.67 (32.15) 26.67 (23.09) 20.00
NiTi 150.00 (52.92) 156.70 (30.55) ⫺6.70
so at the 1.5-mm UDP. Models were ranked from 1 to
27 CuNiTi 103.30 (20.82) 96.67 (15.28) 6.63
5, and models 3 and 5 consistently gained the 2 highest 35 CuNiTi 23.33 (15.28) 16.67 (11.55) 6.66
ranks at both the 1.5- and 1.0-mm UDPs. Model 4 Reflex 120.00 (0.00) 83.33 (5.77) 36.67
displayed the highest ranked ELDP, but 1 of the lowest Tensic 110.00 (10.00) 70.00 (0.00) 40.00
plateau gap values. Models 2 and 4 gave the lowest Nitinol HA 83.33 (15.28) 70.00 (10.00) 13.33
ranked plateau gap for all wires and models 3 and 5 the Model 3
Twistflex 80.00 (10.00) 56.67 (5.77) 23.33
highest. An example of the effect of different models NiTi 220.00 (0.00) 180.00 (10.00) 40.00
on the load deflection graphs for 27 CuNiTi is shown in 27 CuNiTi 163.30 (5.77) 143.30 (5.77) 20.00
Figure 7. These display the difference in force level, 35 CuNiTi 103.30 (11.55) 83.33 (5.77) 19.97
plateau, and hysteresis for the same wire with 3 Reflex 150.00 (0.00) 100.00 (10.00) 50.00
different models. Tensic 150.00 (10.00) 106.70 (5.77) 43.30
Nitinol HA 146.70 (15.28) 110.00 (10.00) 36.70
Model 4
Effect of temperature variation
Twistflex 60.00 (26.46) 6.67 (5.77) 53.33
At 35.5°C and 44.0°C, the superelastic nickel- NiTi 130.00 (10.00) 113.30 (15.28) 16.70
titanium wires displayed a superelastic plateau of vary- 27 CuNiTi 60.00 (20.00) 63.33 (23.09) ⫺3.33
35 CuNiTi 0 (0) 0 (0) 0
ing extent that was rarely present at 22.0°C. ANOVA
Reflex 73.33 (5.77) 56.67 (5.77) 16.66
results indicated that temperature variation had a sta- Tensic 16.67 (15.28) 3.33 (5.77) 13.34
tistically significant effect on values at the P ⬍ .05 Nitinol HA 36.67 (15.28) 43.33 (15.28) ⫺6.66
level. Table VII presents the results of the post hoc Model 5
analysis evaluating the differences between the 7 wires Twistflex 143.00 (5.77) 93.33 (5.77) 49.67
NiTi 256.70 (11.55) 196.70 (5.77) 60.00
at the 3 temperatures. As expected, the load values for
27 CuNiTi 166.70 (11.55) 136.70 (5.77) 30.00
Twistflex had little variation when the temperature was 35 CuNiTi 103.30 (5.77) 73.33 (5.77) 29.97
increased or decreased, with values remaining inside 1 Reflex 133.30 (11.55) 93.33 (5.77) 39.97
standard deviation. However, the NiTi wire load values Tensic 136.70 (5.77) 90.00 (0) 46.70
were significantly affected by temperature changes Nitinol HA 156.70 (5.77) 116.70 (5.77) 40.00
despite remaining the highest ranked wire for all UDP, Unloading deflection point; SD, standard deviation.
temperatures at both 1.5- and 1.0-mm UDPs.
For the 5 HASN wires, temperature change had a
significant effect on ELDP values and load values at change on the load-deflection curves for Nitinol HA in
both UDPs. Compared with the 35.5°C load values, the model 5.
values at 22.0°C were at least halved and, at 44.0°C,
were 1.5 to 2 times larger. This meant that at 44.0°C, Effect of varying wire deflection
the higher ranked 27 CuNiTi and Nitinol HA (3M Table VIII presents the mean values for the 4
Unitek, Monrovia, Calif) delivered forces approaching different load-deflection tests (1.0-4.0 mm) including
those of NiTi at 35.5°C at the same deflection. The the load values, ELDPs, and plateau gap. Nickel-
ELDP became negligible (⬍0.1 mm) for all 5 wires at titanium-based wires rarely displayed a horizontal pla-
44.0°C but increased to a range of 0.4 to 1.4 mm at teau of any duration in the 1.0 mm deflection tests (Fig
22.0°C. Figure 8 depicts the effect of temperature 9). Figure 10 presents rankings of the wire load values
488 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

Table V. Ranking of wires during unloading from 2.0 mm deflection at 35.5°C

Load at 1.5-mm UDP Load at 1.0-mm UDP ELDP Plateau gap

Mean Rank Mean Rank Mean Rank Mean Rank


Wire (g) (of 7) (g) (of 7) (mm) (of 7) (g) (of 7)

Twistflex 73.33 2 41.33 1⫽ 0.3353 5⫽ 32.00 5⫽


NiTi 177.30 7 151.30 7 0 1⫽ 26.00 4*
27 CuNiTi 113.30 5⫽ 102.00 6 0 1⫽ 11.33 1⫽
35 CuNiTi 54.67 1 40.67 1⫽ 0.3180 5⫽ 14.00 1⫽*
Reflex 109.30 5⫽ 76.67 4⫽ 0.0220 1⫽ 32.67 5⫽
Tensic 94.67 3⫽ 63.33 3 0.3013 5⫽ 31.33 5⫽
Nitinol HA 92.67 3⫽ 72.00 4⫽ 0 1⫽ 20.67 3*
SD 14.89 10.51 0.1453 9.76

*Not statistically different from next higher ranked wire.


UDP, Unloading deflection point; ELDP, end load deflection point; SD, standard deviation.

Table VI. Ranking of model designs during unloading from 2.0-mm deflection at deflection at 35.5°C

Load at 1.5-mm UDP Load at 1.0-mm UDP ELDP Plateau gap

Model Mean Rank Mean Rank Mean Rank Mean Rank


Design (g) (of 5) (g) (of 5) (mm) (of 5) (g) (of 5)

1 64.76 2 50.00 2 0 1⫽ 14.76 1⫽


2 90.95 3 74.29 3 0.1671 4 16.67 1⫽
3 144.80 4 111.40 4⫽ 0 1⫽ 33.33 4
4 53.81 1 40.95 1 0.5305 5 12.86 1⫽
5 156.70 5 114.30 4⫽ 0.0000 1⫽ 42.38 5
SD 14.89 10.51 0.1453 9.76

UDP, Unloading deflection point; ELDP, end load deflection point; SD, standard deviation.

for the 4 test deflections at the initial UDP and final Of the 5 HASN wires, 27 CuNiTi provided 1 of the
UDP, respectively. highest-ranked unloading values for the 1-mm and
The 35 CuNiTi wire was consistently 1 of the 2-mm load-deflection tests; however, at the higher
lowest ranked wires at all UDPs in each load-deflection load-deflection tests (3 and 4 mm), it ranked low at the
test (1.0-4.0 mm). It also had 1 of the highest ranked initial UDP measurement but ranked high at the final
ELDPs in the 2-mm, 3-mm, and 4-mm load-deflection 1-mm UDP. This, combined with the lowest ranked
tests, ranging from 0.3 to 1.5 mm. This suggests that plateau gap, indicates that 27 CuNiTi behaved very
when this wire is deflected above 1.0 mm, it will not superelastically, especially at 3 and 4 mm, because it
provide a force below approximately a third of the delivered a low constant force over a low gradient
maximum deflection value when unloading. plateau span. Reflex, Tensic, and Nitinol had compara-
Twistflex, although having higher-ranked plateau ble ranks for most of the tests, and their rankings
gap and ELDP, provided low-ranked unloading values,
typically ranged between those for 35 CuNiTi and
comparable with some HASN wires in all deflection
NiTi.
tests.
Although load values vary according to the differ-
NiTi wire consistently gave the highest-ranked
unloading values at both UDPs in each load-deflection ent deflection tests, the ranking for models was very
test. It also provided 1 of the lowest-ranked ELDPs, similar to those presented previously for 2-mm deflec-
with load being provided at all unloading deflections in tion (Table VI), with models 3 and 5 having the 2
the 1-mm and 2-mm load-deflection tests, and produced highest-ranked load values at most deflections. The
1 of the lowest plateau gaps in the 3-mm and 4-mm ELDPs for models 3 and 5 were negligible in all
load-deflection tests. Because NiTi wire provided the deflection tests, in contrast with model 4, which gave
highest unloading values, this plateau gap represented a no load below a mean of 3.5 mm in the 4-mm
particularly small proportional change in force level. load-deflection test. Figure 11 shows the different
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 489
Volume 121, Number 5

Fig 7. Load-deflection graphs.

effect of model designs on the load-deflection graphs of Fig 8. Load-deflection graphs.


the Tensic wire.
the differing results arose because of interbatch varia-
DISCUSSION tion and the small difference in test temperature.30
Raw data from mechanical testing of wires should Similarly, despite the attempt in the present study to
be interpreted with caution. This is emphasized by the design models resembling clinical conditions, there is
variation in results obtained for each wire depending on questionable validity in extrapolating numerical load
the model design. Test conditions should therefore be values to clinical situations, all of which tend to present
identical before valid comparisons of such data can be unique circumstances. Therefore, in accordance with
made between studies. This point is illustrated by the the suggestion of Dowling et al,23 the present study
marked difference in load values between those of the placed greater emphasis on comparative rankings of
present study and those reported by Nakano et al.4 for wire and model performance than on actual data. It is
3-point bending of 27 CuNiTi and 35 CuNiTi. Al- evident from Table IV that the ranking of load values of
though test conditions differed only in test span (by 1.5 the various wires differed according to the model used.
mm) and temperature (by 1.5°C), the load values Three-point bending tests did not consistently represent
reported by Nakano et al4 were less than half those of other model designs, indicating that their status as the
the present study. The wire manufacturer suggested that standard wire test may not be entirely appropriate.
490 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

Table VII. Ranking of wires during unloading from 2.0-mm deflection at 22.0°C, 35.5°C, and 44.0°C

Load (g) at 1.5-mm UDP Load (g) at 1.0-mm UDP

22.0°C 35.5°C 44.0°C 22.0° 35.5°C 44.0°C

Wire Mean Rank Mean Rank Mean Rank Mean Rank Mean Rank Mean Rank

Twistflex 82.00 6 73.33 2 72.00 1 42.67 6 41.33 1⫽ 48.00 1


NiTi 109.30 7 177.30 7 221.30 7 89.33 7 151.30 7 186.00 7
27 CuNiTi 44.67 3⫽ 113.30 5⫽ 168.70 6 26.67 5 102.00 6 148.00 6
35 CuNiTi 14.00 1 54.67 1 112.00 2 0 1⫽ 40.67 1⫽ 98.00 2⫽
Reflex 53.33 3⫽ 109.3 5⫽ 146.70 3⫽ 18.00 3⫽ 76.67 4⫽ 102.70 2⫽
Tensic 34.00 2 94.67 3⫽ 141.30 3⫽ 5.33 1⫽ 63.33 3 102.70 2⫽
Nitinol HA 45.33 3⫽ 92.67 3⫽ 151.30 3⫽ 19.33 3⫽ 72.00 4⫽ 128.00 5
SD 11.54 14.89 14.00 8.16 10.51 12.35

UDP, Unloading deflection point; ELDP, end load deflection point.


*Not statistically different from next higher ranked wire.

Fig 10. A, Ranking for wire load values at initial unload-


ing deflection point; B, ranking for wire load values at
Fig 9. Load-deflection graphs. final unloading deflection point.
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 491
Volume 121, Number 5

ELDP Plateau gap (g)

22.0°C 35.5°C 44.0°C 22.0° 35.5°C 44.0°C

Mean Rank Mean Rank Mean Rank Mean Rank Mean Rank Mean Rank

0.2260 2 0.3353 5⫽ 0.0807 1⫽ 39.33 6⫽ 32.00 5⫽ 24.00 3⫽


0 1 0 1⫽ 0.0580 1⫽ 20.00 3 26.00 4* 35.33 5⫽
0.4227 3 0 1⫽ 0.0527 1⫽ 18.00 1⫽ 11.33 1⫽ 20.67 1⫽
1.4280 7 0.3180 5⫽ 0.0093 1⫽ 14.00 1⫽* 14.00 1⫽* 14.00 1⫽*
0.8147 5 0.0220 1⫽ 0.0180 1⫽ 35.33 6⫽ 32.67 5⫽ 44.00 5⫽
1.0940 6 0.3013 5⫽ 0.0333 1⫽ 28.67 4⫽ 31.33 5⫽ 38.67 5⫽
0.6327 4 0 1⫽ 0.0393 1⫽ 26.00 4⫽ 20.67 3* 23.33 3⫽
0.2224 0.1453 0.0858 7.62 9.76 11.95

Model design Temperature variation


Model rankings showed that the full maxillary arch All 6 nickel-titanium wires examined, including the
with Mini-Diamond brackets (model 4) gave the lowest stabilized austenitic NiTi that is reported not to expe-
load values, suggesting that greater friction between rience shape memory at intraoral temperatures,13 dem-
these brackets and the wires reduced the load values. onstrated temperature sensitivity. The unloading values
By contrast, model 5, comprising the same configura- for the latter wire increased significantly at higher
tion except having the self-ligating Twin-Lock brack- temperatures and decreased at lower temperatures, an
ets, gave the highest average load values for the wires, observation previously made by Tonner and Waters5
nearly 3 times the values for model 4. This suggests and Filleul and Jordan.20 Presumably, at the increased
that Twin-Lock brackets produced significantly less temperature of 44.0°C, all nickel-titanium wires were
friction, and the wires were therefore able to exert in the austenitic phase, which also accounts for their
higher loads. This interpretation is supported by the ELDPs decreasing to a negligible level (⬍0.1 mm) at
work of Pizzoni et al,31 who noted that self-ligating this temperature. The converse occurred at 22.0°C,
brackets provided markedly lower friction than do when the wires were likely to be in the martensitic
conventional brackets. The clinical relevance of friction phase, with a subsequent decrease in load values and an
increase in ELDP. This agrees with the findings of
is, however, difficult to quantify; it is possible that
Tonner and Waters,5 who reported that when the wire
factors influencing bind and release of the wire in
exists solely in the austenitic state a greater stress is
occlusal function could be more important in determin-
required to produce deformation than when other
ing sliding behavior.32
phases coexist, such as the R and the martensite phases
Plateau values were measured in the present study
at lower temperatures. The stages and amounts of these
because this4 and other gradient analyses4,33-35 are
phase transformations cannot be confirmed unless x-ray
frequently used to express a measure of superelasticity diffraction (XRD) or differential scanning calorimetry
of nickel-titanium wires. However, the plateau gap in (DSC) techniques are used. The manufacturer of Cu-
isolation may be a limited measure of superelastic NiTi wires used in the present study advises30 that the
behavior. For example, in 3-point bending, the plateau austenitic finish temperature of 35 CuNiTi is 35°C ⫾ 2;
gap was the same (20 g) for NiTi, Reflex, and Nitinol this concurs with Filleul and Jordan,20 who, using DSC,
HA, suggesting that these 3 wires behaved similarly. found that the austenitic finish of the 35 CuNiTi wire
However, as a proportion of the load value at the 1.5- used in their study was 36.6°C. Force variation is
mm UDP, the 20-g plateau gap represented a 50% drop particularly significant at temperatures close to the
in force for Nitinol HA but only a 15.38% drop for TTR, which possibly contributes to the differing results
NiTi. This indicates that although wires may have in 3-point bending between the current study and that of
comparable plateau gradients, consideration should Nakano et al.4
also be given to the load levels associated with these Research by Meling and Ødegaard36 and Barwart21
plateau measurements. has also shown that the behavior of HASN wires is
492 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

Table VIII. Ranking of wires during unloading from 1.0- to 4.0-mm deflections at 35.5°C

1.0-mm deflection

Load at 1.0-mm UDP Load at 0.5-mm UDP ELDP Plateau gap

Wire type Mean (g) Rank Mean (g) Rank Mean (mm) Rank Mean (g) Rank

Twistflex 117.30 1 28.67 1 0.0653 7 88.67 1


NiTi 311.30 7 114.00 7 0 1⫽ 197.30 7
27 CuNiTi 247.30 6 90.67 6 0 1⫽ 156.70 2⫽
35 CuNiTi 200.70 2⫽ 46.00 2 0.0207 1⫽* 154.70 2⫽
Reflex 216.00 4⫽ 70.67 3⫽* 0 1⫽ 145.30 2⫽
Tensic 212.00 2⫽* 62.00 3⫽ 0.0500 6 150.00 2⫽
Nitinol HA 224.00 4⫽ 74.67 5 0 1⫽ 149.30 2⫽
SD 19.88 14.53 0.0475 16.62

3.0-mm deflection

Load at 2.5-mm UDP Load at 1.0-mm UDP ELDP Plateau gap

Wire type Mean (g) Rank Mean (g) Rank Mean (mm) Rank Mean (g) Rank

Twistflex 86.00 2⫽* 34.00 1⫽ 0.7140 6 52.00 5⫽


NiTi 149.30 7 125.30 7 0.3060 1⫽ 24.00 2⫽
27 CuNiTi 90.67 2⫽* 82.67 6 0.5093 3⫽* 8.00 1
35 CuNiTi 54.67 1 34.67 1⫽ 0.8493 7 20.00 2⫽
Reflex 105.30 6 60.00 3⫽ 0.1760 1⫽* 45.33 5⫽
Tensic 98.00 5* 53.33 3⫽ 0.5927 3⫽* 44.67 5⫽
Nitinol HA 79.33 2⫽ 54.67 3⫽ 0.5113 3⫽* 24.67 2⫽
SD 17.02 13.35 0.4062 14.21

UDP, Unloading deflection point; ELDP, end load deflection point.


*Not statistically different from next higher ranked wire.

highly temperature dependent, and performance may be flection up to 4 mm. The clinical impression that HASN
influenced by environmental temperature fluctuations. wires are superior to NiTi and nonsuperelastic arch-
The temperatures in the current study were selected wires may be because they are more readily engaged
from studies by Moore et al,37 Meling22 and Filleul and into such grossly displaced teeth with their temperature
Jordan20 to represent a range of intraoral temperatures. phase transformation. In the present study, 27 CuNiTi
Moore et al37 observed that an average young adult could be claimed to have demonstrated superior super-
male’s maxillary archwire at the labial surface of the elastic behavior over the 4 deflection tests by providing
central incisors could be expected, over a 24-hour a reasonably constant low load-deflection for a long
period, to be in the range of 33.0°C to 37.0°C for 79% range of deflections, but the magnitude of the optimal
of the time, below it for 20% of the time, and above it unloading force during initial alignment remains con-
for only 1% of the time. Many studies3,4,20,37 have used
troversial.18,38,39
37°C to represent intraoral temperature, but Moore et
Analysis of the unloading curves in the present
al37 concluded that 35.5°C would be more appropriate
study displayed a variation in plateau gradient and
for in vitro testing of orthodontic wires. This factor
ELDPs of individual wires deflected to different de-
becomes increasingly important when examining the
performance of the HASN archwires. A wire with a grees. The ELDPs for all wires increased at 3- and
temperature transition range below 35°C (eg, 27 Cu- 4-mm deflections, and, in the 4-mm test, all had
NiTi) should be fully active for a considerably longer finished unloading by 0.8 mm. This implies, when 1 of
period than those manufactured to be fully active at these wires is displaced to 4 mm to engage a malposi-
37°C to 40°C. tioned tooth, that it will not provide a force beneath 0.8
mm of unloading. However, it is possible that, clini-
Wire deflection cally, cyclic loading from masticatory forces may
Full-bracket engagement during initial alignment in contribute to a bind-and-release function of the
grossly irregular dentitions may involve archwire de- wire,40,41 so that a force is generated beneath this
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 493
Volume 121, Number 5

2.0-mm deflection

Load at 1.5-mm UDP Load at 1.0-mm UDP ELDP Plateau gap

Mean (g) Rank Mean (g) Rank Mean (mm) Rank Mean (g) Rank

73.33 2 41.33 1⫽ 0.3353 5⫽ 32.00 5⫽


177.30 7 151.30 7 0 1⫽ 26.00 4*
113.30 5⫽ 102.00 6 0 1⫽ 11.33 1⫽
54.67 1 40.67 1⫽ 0.3180 5⫽ 14.00 1⫽*
109.30 5⫽ 76.67 4⫽ 0.0220 1⫽ 32.67 5⫽
94.67 3⫽ 63.33 3 0.3013 5⫽ 31.33 5⫽
92.67 3⫽ 72.00 4⫽ 0 1⫽ 20.67 3*
14.89 10.51 0.1453 9.76

4.0-mm deflection

Load at 3.5-mm UDP Load at 1.0-mm UDP ELDP Plateau gap

Mean (g) Rank Mean (g) Rank Mean (mm) Rank Mean (g) Rank

74.67 2⫽ 21.33 1⫽ 1.1247 1⫽* 53.33 7


116.00 7 101.30 7 0.9367 1⫽ 14.67 2⫽
77.33 2⫽* 74.00 6 0.9287 1⫽ 3.33 1
47.33 1 30.00 1⫽ 1.4740 7 17.33 2⫽
96.67 6 54.00 3⫽ 0.7867 1⫽ 42.67 5⫽
89.33 5* 46.67 3⫽ 0.8340 1⫽ 42.67 5⫽
71.33 2⫽ 49.33 3⫽ 0.9093 1⫽ 22.00 2⫽
18.31 19.30 0.6181 13.87

ELDP. This masticatory force, however, may also design of the test model, including bracket type, and
cause a permanent set in the wire.42 the amount of deflection.
Although statistically there may be a difference in 2. The NiTi wire provided the highest unloading val-
the performance of individual wires from various me- ues for every test deflection and model design.
chanical test simulations, this does not necessarily 3. The rankings of the remaining wires were altered by
indicate that differences will exist in clinical perfor- test conditions, with Twistflex providing compara-
mance. West et al15 compared multistrand round stain- ble values with many HASN wires.
less steel wire with superelastic NiTi in a randomized 4. The plateau gap was considered to be a limited
clinical trial to evaluate tooth alignment ability and measure of wire performance in isolation.
found that the degree of initial alignment achieved with 5. The 6 superelastic nickel-titanium wires all dis-
the 2 wires was similar over a 6-week period. In a played temperature sensitivity at the 3 test temper-
crowded dentition, the high forces are possibly dissi- atures, resulting in significant effects on unloading
pated through interdental contacts and in overcoming values and ELDPs.
friction between the bracket, wire, and associated 6. Comparisons of 3-point bending results with those
ligatures.18,41,43 A recent study by Evans et al16 com- of a similar recent study suggest possible interbatch
pared the alignment performance of multistrand stain- variation of some of these wires.
less steel, superelastic NiTi, and HASN over a longer 7. The Twin-Lock self-ligating brackets apparently
period in a controlled randomized clinical trial and produce less friction and allow a higher force to be
found no significant difference in aligning capability expressed; however, important aspects such as bio-
among the 3 wires. logical resistance and bracket angulation44 were not
addressed by this study and need further investiga-
CONCLUSIONS tion.
1. This study shows that the load-deflection perfor- 8. Wire rankings from 3-point bending tests were not
mance of 0.016-in alignment wires depends on the consistently representative of those with other
494 Wilkinson et al American Journal of Orthodontics and Dentofacial Orthopedics
May 2002

2. Oltjen JM, Duncanson MG, Ghosh J, Nanda RS, Currier GF.


Stiffness-deflection behaviour of selected orthodontic wires.
Angle Orthod 1997;67:209-18.
3. Mullins WS, Bagby MD, Norman TL. Mechanical behaviour of
thermo-responsive orthodontic archwires. Dent Mat 1996;12:
308-14.
4. Nakano H, Satoh S, Norris R, Jin T, Kamegai T, Ishikawa F, et
al. Mechanical properties of several nickel-titanium alloy wires
in three-point bending tests. Am J Orthod Dentofacial Orthop
1999;115:390-5.
5. Tonner RIM, Waters NE. The characteristics of super-elastic
NiTi wires in three-point bending. Part 1: The effect of temper-
ature. Eur J Orthod 1994;16:409-19.
6. Tonner RIM, Waters NE. The characteristics of super-elastic
NiTi wires in three-point bending. Part 2: Intra-batch variation.
Eur J Orthod 1994;16:421-5.
7. Khier SE, Brantley WA, Fournelle RA. Bending properties of
superelastic and nonsuperelastic nickel-titanium orthodontic
wires. Am J Orthod Dentofacial Orthop 1991;99:310-8.
8. Hudgins JJ, Bagby MD, Erickson LE. The effect of long-term
deflection on permanent deformation of nickel-titanium arch-
wires. Angle Orthod 1990;60:283-8.
9. Burstone CJ, Morton JY. Chinese NiTi wire—a new orthodontic
alloy. Am J Orthod Dentofacial Orthop 1985;87:445-52.
10. Thayer TA, Bagby MD, Moore RN, DeAngelis RJ. X-ray
diffraction of nitinol orthodontic arch wires. Am J Orthod
Dentofacial Orthop 1995;107:604-12.
11. Kusy RP. A review of contemporary archwires: their properties
and characteristics. Angle Orthod 1997;67:197-208.
12. Wayman CM, Duerig TW. Engineering aspects of shape memory
alloys. London: Butterworth-Heinemann; 1990.
13. Hurst CL, Duncanson MG, Nanda RS, Angolkar PV. An evalu-
ation of the shape-memory phenomenon of nickel-titanium
orthodontic wires. Am J Orthod Dentofacial Orthop 1990;98:
72-6.
14. Kapila S, Sachdeva R. Mechanical properties and clinical appli-
cations of orthodontic wires. Am J Orthod Dentofacial Orthop
1989;96:100-9.
15. West AE, Jones ML, Newcombe RG. Multiflex versus super-
elastic: a randomized clinical trial of the tooth alignment ability
of initial arch wires. Am J Orthod Dentofacial Orthop 1995;108:
464-71.
16. Evans TJ, Jones ML, Newcombe RG. Clinical comparison and
performance perspective of three aligning arch wires. Am J
Orthod Dentofacial Orthop 1998;114:32-9.
Fig 11. Load-deflection graphs. 17. O’Brien K, Lewis D, Shaw W, Combe E. A clinical trial of
aligning archwires. Eur J Orthod 1990;12:380-4.
model designs; this supports the suggestion that this 18. Waters NE, Stephans CD, Houston WJB. Physical characteristics
test may have limited applicability to clinical con- of orthodontic wires and archwires—part 1. Br J Orthod 1975;
2:15-24.
ditions.
19. Rock WP, Wilson HJ. Forces exerted by orthodontic aligning
9. Until the universal adoption of improved standard- archwires. Br J Orthod 1988;15:255-9.
ized and reproducible test conditions for wires and 20. Filleul MP, Jordan L. Torsional properties of Ni-Ti and copper
uniform measurement selection, the ranking of re- Ni-Ti wires: the effect of temperature on physical properties. Eur
sults, as used in the present study, may be the most J Orthod 1997;19:637-46.
appropriate method of analysis. 21. Barwart O. The effect of temperature change on the load value of
Japanese NiTi coil springs in the superelastic range. Am J Orthod
Dentofacial Orthop 1996;110:553-8.
REFERENCES 22. Meling TR. The effect of short-term temperature changes on the
1. Miura F, Mogi M, Ohura Y, Hamanaka H. The super-elastic mechanical properties of rectangular nickel titanium archwires
property of Japanese NiTi alloy wire for use in orthodontics. Am J tested in torsion. Angle Orthod 1998;68:369-76.
Orthod Dentofacial Orthop 1986;90:1-10. 23. Dowling PA, Jones WB, Lagerstrom L, Sandham JA. An
American Journal of Orthodontics and Dentofacial Orthopedics Wilkinson et al 495
Volume 121, Number 5

investigation into the behavioural characteristics of orthodontic relevance to orthodontic treatment. Eur J Orthod 1995;17:395-
elastomeric modules. Br J Orthod 1998;25:197-202. 402.
24. Tidy DC. Frictional forces in fixed appliances. Am J Orthod 34. Meling T, Ødegaard J, Holthe K, Meling EO, Segner D. A
Dentofacial Orthop 1989;96:249-54. formula for the displacement of an arch wire when subjected to
25. Ogata RH, Nanda R, Duncanson MG, Sinha PK, Currie GF. a second-order couple. Am J Orthod Dentofacial Orthop 1998;
Frictional resistances in stainless steel bracket-wire combinations 113:632-40.
with effects of vertical deflections. Am J Orthod Dentofacial 35. Chang R, Nikolai RL. Temperature influences on nickel-titani-
Orthop 1996;109:535-42. um-alloy-wire responses in flexure. J Dent Res 1994;73:ADR
26. Pratten DH, Popli K, Germane N, Gunsolley J. Frictional Abstracts No. 1767.
resistance of ceramic and stainless steel orthodontic brackets. 36. Meling TR, Ødegaard J. The effect of temperature on the elastic
Am J Orthod Dentofacial Orthop 1990;98:398-403. responses to longitudinal torsion of rectangular nickel-titanium
27. Bednar JR, Grueneman GW, Sandrik JL. A comparative study of archwires. Angle Orthod 1998;68:357-68.
37. Moore RJ, Watts JTF, Hood JAA, Burritt DJ. Intra-oral temper-
frictional forces between orthodontic brackets and arch wires.
ature variation over 24 hours. Eur J Orthod 1999;21:1-13.
Am J Orthod Dentofacial Orthop 1991;100:513-22.
38. Quin R, Yoshikawa DA. A reassessment of force magnitude in
28. Moyers RE, van der Linden FPGM, Riolo ML, McNamara JA.
orthodontics. Am J Orthod 1985;88:252-60.
Standards of human occlusal development. Ann Arbor: Center
39. Lee BW. The force requirements for tooth movement. Part 1:
for Human Growth and Development; University of Michigan;
Tipping and bodily movement. Aust Orthod J 1995;13:238-48.
1976.
40. Thurow RC. Edgewise orthodontics. St. Louis: Mosby; 1982.
29. Schaus JG, Nikolai RJ. Localized, transverse, flexural stiffnesses 41. Read-Ward GE, Jones SP, Davies EH. A comparison of self-
of continuous arch wires. Am J Orthod Dentofacial Orthop ligating and conventional orthodontic bracket systems. Br J
1986;89:407-14. Orthod 1997;24:309-17.
30. Farzin-Nia F. Personal communications. Technical Department, 42. Drescher D, Bourauel C, Schumacher HA. Frictional forces
Materials Research and Development, Ormco Corp; 2000. between bracket and arch wire. Am J Orthod Dentofacial Orthop
31. Pizzoni L, Ravnholt G, Melson B. Frictional forces related to 1989;96:397-403.
self-ligating brackets. Eur J Orthod 1998;20:283-91. 43. Andreasen GF, Quevedo FR. Evaluation of friction forces in the
32. Kusy RP, Whitley JO. Influence of archwire and bracket dimen- 0.022 ⫻ 0.028 edgewise bracket in vitro. J Biomech 1970;3:151-
sions on sliding mechanics: derivations and determinations of the 60.
critical contact angles for binding. Eur J Orthod 1999;21:199- 43. Articolo LC, Kusy RP. Influence of angulation on the resistance
208. to sliding in fixed appliances. Am J Orthod Dentofacial Orthop
33. Segner D, Ibe D. Properties of superelastic wires and their 1999;115:39-51.

You might also like