Anda di halaman 1dari 17

Periodontology 2000, Vol.

55, 2011, 189204


Printed in Singapore. All rights reserved

 2011 John Wiley & Sons A/S

PERIODONTOLOGY 2000

Calculus-detection technologies
and their clinical application
GRIT MEISSNER & THOMAS KOCHER

Subgingival calculus surfaces are usually covered


with a layer of unmineralized and metabolically
active bacteria. The essential component of conventional periodontal therapy is the effective removal of
these bacterial deposits from the root surface, along
with calculus deposits, in order to create a biologically compatible root surface (10, 45, 63).
While numerous clinical studies have documented
the beneficial effects of complete removal of subgingival calculus on the resolution of inflammation
(11, 45, 63), others have found that gingival tissues
adjacent to root surfaces covered with small polished
calculus spots might have a tendency to heal that is
similar to tissues adjacent to thoroughly cleaned,
calculus-free root surfaces (26, 44). Nevertheless,
periodontal destruction is clearly related to the very
presence of calculus, which may extend the range of
damage associated with plaque microorganisms (36,
61, 64).
Calculus is a porous substance that can adsorb a
variety of toxic products and retain significant levels
of endotoxin, which itself can damage tissue (64).
These toxins are located on, not within, periodontally
diseased root surfaces (7, 22, 43). It was therefore
deduced that extensive removal of cementum is not
necessary, and root surfaces should be treated
carefully during periodontal therapy in order to
selectively remove subgingival calculus and biofilm
without removing the underlying cementum.
Subgingival root debridement currently comprises
the systematic treatment of all diseased root surfaces
using hand-sonic and or ultrasonic instruments,
followed by tactile control with a periodontal probe,
explorer or curette, until the root surface feels
smooth and clean. However, traditional tactile perception of the subgingival environment without visible access before and after treatment frequently
lacks sensitivity, specificity and reproducibility, and

thus may lead to the unwanted removal of cementum, residual calculus, or both (6, 25, 27, 47, 57).
Clinicians are often uncertain about the nature of a
subgingival root surface while performing periodontal instrumentation. The correct evaluation of a
cleaned surface is key to enable thorough and substance-sparing debridement. To support the clinicians decision to either stop or continue therapy, the
past few years have witnessed the development of
several calculus-detection techniques based on different technologies. Current technologies for calculus
identification include detection-only systems (a
miniaturized endoscope, a device based on light
reflection and a laser that activates the tooth surface to
fluoresce) as well as combined calculus-detection and
calculus-removal systems [an ultrasonic oscillationbased system that analyzes impulses reflected from
the tooth surface, and a system combining erbiumdoped yttrium aluminium garnet (Er:YAG) and diode
lasers] (Tables 1 and 2). The aim of this article was to
provide a critical review of these devices based on
currently available clinical and experimental data.

Detection-only systems
Fiberoptic endoscopy-based technology
The idea to modify a medical endoscope for periodontal use has, to date, been realized in only one
device (Perioscopy; Perioscopy Inc., Oakland, CA,
USA), which was introduced in the year 2000. Perioscopy is a minimally invasive miniature periodontal
endoscope which is inserted into the periodontal
pocket and permits visualization of the root surface
within the subgingival environment at magnifications
of 2448 (Fig. 1). The system consists of a 1 mm,
10,000-pixel fiberoptic bundle surrounded by multiple

189

Meissner & Kocher

Table 1. Automated calculus-detection technologies


Treatment goal
Calculus detection only

Combined calculus detection and


removal

Technology

Device name

Fiberoptic endoscopy

Perioscopy

Spectro-optical technology

Detectar

Autofluorescence

Diagnodent

Ultrasound

Perioscan

Laser and autofluorescence

Keylaser3

illumination fibers, a light source, an irrigation system and a liquid crystal display monitor. Clinicians
can observe the subgingival root surface, tooth
structure and residual calculus in real time. The
magnified images can be viewed on the monitor in
real time, and images and videos can be captured and
saved in computer files. The endoscope may help to
identify, locate and treat calculus spots during
instrumentation of residual calculus at the time of, or
after, scaling. To be proficient in the endoscopic
technique a training period of at least 8 h is necessary
to learn the procedure and practical experience is
required for up to 4 weeks subsequently (59, 60).
In the first clinical study, nonresponding periodontal sites (n = 44; probing depth 58 mm) were
treated by subgingival root debridement with or
without use of the dental endoscope (5). No significant changes regarding pocket depth reduction were
reported in either group, 1 and 3 months after
treatment, compared with baseline. Moreover, the
gingival crevicular fluid flow rate, prostaglandin E2
and interleukin-1beta levels decreased without
showing significant differences between the groups.
Additionally, a rather long treatment time, of 45 min
per experimental site, was noted for the Perioscopy
procedure.
In a study evaluating the histologic response to the
removal of calculus and biofilm with the aid of the
dental endoscope (65), a total of 12 teeth from six
patients were extracted 6 months after endoscopeaided scaling and root planing. Histological evidence
showed formation of a long junctional epithelium,
bone repair and no signs of chronic inflammation.
However, a control group that received scaling and
root planing alone was not included and therefore
the incremental effect attributable to the use of the
endoscope was not determined.
A randomized, controlled, clinical study evaluated
the percentage of residual calculus after tooth
extraction (20) in 100 single-rooted teeth of 15
patients. The teeth were treated by hand- and ultra-

190

sonic instruments until the root surface was found to


be clean, as assessed by either an explorer or the
periodontal endoscope. After extraction, a higher
percentage of residual calculus covering the root
surface was detected microscopically in the explorer
group than in the endoscope group (D = 2.1%). The
difference was statistically significant only in deeper
pockets and in interproximal sites (pocket depth
> 6 mm; D = 2.9%) compared with buccal sites
(pocket depth > 4 mm; D = 1.3%). A correlation was
found between shorter treatment time and increasing
experience of the operator for treatment with the
endoscope, a finding confirmed by a companion
study (41). However, the treatment results of the
latter study showed some discrepancies. Out of 24
patients, a total of 70 molars were treated in vivo
either by scaling and root planing only or by scaling
and root planing plus dental endoscopy, followed by
extraction. Overall,1.2% less residual calculus covering the root surface was found in the endoscopy
group (12.3%) compared with the scaling and root
planing group (13,5%). No differences in residual
calculus were found in deep pockets, furcation areas
or on buccal lingual surfaces. Only interproximal
pockets with a depth of < 6 mm had significantly less
residual calculus in the endoscope group compared
with the scaling and root planing group. Thus, at least
for multi-rooted teeth, the beneficial effect of the
endoscope-aided scaling and root planing remains
questionable.
Taken together, only one clinical study to date has
investigated the clinical effects after the application of
fiberoptic technology. No differences were found
regarding pocket depth reduction between scaling and
root planing alone and endoscope-aided scaling and
root planing. Histologic healing, which was assessed
on extracted teeth 6 months after endoscope-aided
scaling and root planing, was not compared with
scaling and root planing alone in a randomized clinical
study. Microscopic analysis of root surfaces after
endoscopy-aided scaling and root planing showed a

Calculus-detection technologies

Table 2. Studies reviewed in this article


Instrument

Reference

Design

Sample size

Method

Results

Diagnodent

(31)

In vitro study

10 teeth, 271 sites

Fluorescence was measured at five teeth and


reproducibility was
tested (at all five teeth)
Effect of root
debridement on
fluorescence was tested

A clean root surface was


indicated with a median
value of 6.2, in contrast to a
median value of 57.7 on
the root where calculus was
found
Not influenced by the fluid
High reproducibility
Fluorescence values after root
debridement were similar to
those for a clean root surface

Diagnodent

(17)

In vitro study

A total of 30 teeth,
For each medium,
10 teeth were
included

Fluorescence was
measured in medium,
air, saline solution
and blood

Significant differences in
fluorescence between calculus
and cementum in all fluids
Air: cementum, 0.4; calculus,
54.1
Saline solution: cementum,
0.4; calculus, 60.7
Blood: cementum, 2.1;
calculus, 39.6

Diagnodent

(16)

Keylaser 3

(30)

Hand instrumentation Surface area of residual calculus


Multirooted teeth:
with and without
hand instrumentation:
Diagnodent. In total,
0.5 0.48 107 lm2
120 surfaces were
evaluated
Diagnodent: 0.27 0.43 107
lm2 (P = 0.02)
Single-rooted teeth:
hand instrumentation:
0.19 0.37 107 lm2
Diagnodent:
0.11 0.26 107 lm2
(P = 0.19)
Threshold 5 [U]; the median
ERL (140 mJ per pulse,
In vitro study
Twenty teeth
residual calculus was
10 Hz), with a chiselcovered with
11 (078)%
subgingival calculus shaped glass-fiber tip
Threshold 1 [U]; the median
(0.4 1.65 mm); water
were treated with
irrigation (1 ml min) residual calculus was 0 (026)%
an ERL
Fluorescence threshold Laser-treated cementum thicklevel of 5 [U] was reduced ness [median, 80 (0250) mm]
at intervals of 1 [U] for Untreated opposite side [medevery laser treatment ian, 90 (30250) mm] (P < 0.05)

Keylaser 3

(53)

In vitro study

A total of 40 teeth;
20 teeth were
included for each
treatment

Randomized,
single-masked
study

Three teeth per patient Histologically, ERL produced


Twelve patients,
were treated with an ERL homogeneous and smooth root
each with six
surfaces
[ERL1, 100 mJ per pulse;
periodont
ERL2, 120 mJ per pulse; Calculus was almost selectively
ally diseased
single-rooted teeth ERL3, 140 mJ per pulse; removed, no thermal damage,
no cementum loss, mean
10 Hz; water irrigation;
chisel-shaped glass-fiber treatment time needed with
tip (0.4 1.65 mm); the ERL was comparable to that
for hand instrumentation
transmission factor 0.85]
Hand instrumentation resulted
and three teeth per
patient were treated in significantly higher values for
with the Vector system residual calculus and in more
or hand instrumentation, root surface damage than laser
treatment
or were untreated
(control)

191

Meissner & Kocher

Table 2. (Continued)
Instrument

Reference

Design

Sample size

Keylaser 3

(55)

Randomized
clinical study

Twenty-four periodontally diseased


single-rooted teeth

Keylaser 3

(56)

Randomized,
controlled,
split-mouth
study

Keylaser 3

(62)

192

Method

Results

Histologically, calculus was


ERL, water irrigation
[160 mJ per pulse and selectively removed No thermal
damage
chisel-shaped tip
Results obtained following
(1.65 0.5 mm);
calculated energy density treatment with the ERL were
19.4 J cm2 per pulse; comparable to those obtained
by hand instruments
10 Hz] vs. hand
instrumentation
ERL, water irrigation Average treatment time in both
Twenty patients,
groups was 5 min for
single-rooted teeth [160 mJ per pulse; 10 Hz;
single-rooted teeth and 9 min
chisel-shaped tip
[n = 407 for laser
for multirooted teeth
(1.65 0.5 mm); calcutreatment (ERL),
All clinical parameters
lated energy density
n = 383 for UI]
investigated showed improve136 mJ per pulse; or
multirooted teeth
ment in both groups, which was
chisel-shaped tip
(n = 269 for laser
treatment, n = 247 (1.1 0.5 mm); calculated significant between baseline
energy density 114 mJ and 6 months post-treatment
for UI
Bleeding on probing:
per pulse]
ERL: baseline, 40%; 6 months,
17%
UI: baseline, 46%; 6 months,
15%
Clinical attachment level gain:
ERL: after 3 months, 1.48 0.73;
after 6 months, 1.11 0.59
UI: after 3 months, 1.53 0.67;
after 6 months, 1.11 0.46
There were no statistically
significant differences between
the groups

Baseline:
Single masked, Twenty patients at Treatment either by ERL
randomized, recall visit with at [160 mJ per pulse;10 Hz; Mean pocket depth: ERL, 6 mm;
UI, 5.8 mm
water irrigation;
least two residual
controlled,
chisel-shaped tips
pocket depths of
split-mouth
After 1 month significant differdesign study > 5 mm in each jaw (0.5 1.1 mm)] or by a
ences:
piezoelectric ultrasonic Mean pocket depth reduction:
scaler (UI) (Piezon Master
ERL, 0.9 mm; UI, 0.5 mm
400; EMS, Nyon,
(P < 0.05)
Switzerland)
Mean clinical attachment
Clinical and microbiologic level gain: ERL, 0.5 mm; UI,
effects at 1 and 4 months
0.06 mm (P < 0.01)
post-treatment were
After 4 months no significant
evaluated
differences:
Mean pocket depth reduction:
ERL, 1.1 mm; UI, 1.0 mm
Mean clinical attachment level
gain: ERL, 0.6 mm; UI, 0.4 mm
Both treatment modalities
resulted in reduction of subgingival microflora, with no
differences between the groups
The patients preference was
laser instrumentation

Calculus-detection technologies

Table 2. (Continued)
Instrument

Reference

Design

Sample size

Keyaser 3

(13)

Single-blinded,
randomized,
controlled,
specific quadrant design
study

Seventy-two
patients with
periodontal
disease

Perioscopy

(5)

Randomized
patient
matched-site
design study

Perioscopy

(20)

Randomized Fifteen patients, a


clinical and
total of 100 sites,
in vitro study Single-rooted teeth

Perioscopy

(41)

Randomized
clinical and
in vitro study

Method

Results

Treatment per quadrant: All four treatment modalities


resulted in a significant
hand instruments
reduction of Porphyromonas
(Gracey curettes
(Hu Friedy), feedback- gingivalis, Prevotella intermedia,
Tannerella forsythia and
controlled ERL (160 mJ
Treponema denticola after
per pulse;10 Hz; water
3 months. Laser and sonic
irrigation; chisel-shaped
instrumentation failed to
tips of 0.5 1.65 and
reduce Aggregatibacter
0.5 1.1 mm), sonic
actinomycetemcomitans
scaler (SONICflexs system
significantly
LUX 2003 L; KaVo)
The patients preference
and a piezoelectric
was UI
ultrasonic scaler (Piezon
Master 400, EMS)
Bacterial samples were
investigated at baseline,
and at 3 and 6 months
post-treatment

Group A: scaling and root Plaque index, bleeding on


Six patients on
probing, clinical attachment
planing plus explorer
maintenance
level gain: no significant
therapy, 44 sites Group B: scaling and root
differences after 3 months
with pocket depth planing plus Perioscopy
between the groups
Treatment until root
58 mm
Pocket depth: decrease of
surface was considered
2 mm in both groups, no
to be clean
significant differences
Evaluation of plaque
index, bleeding on prob- Treatment duration unrealistic
for clinical use
ing, clinical attachment
level after 3 months
Group A: scaling and 2.1% more residual calculus in
root planing plus
the explorer group
explorer
Statistical significance only in
Group B: scaling and
interproximal sites (pocket
root planing plus
depth > 6 mm; 2.9%)
Perioscopy
Treatment duration: endoscope
Treatment until root
group showed a significant
surface was considered to
decrease of time with
increasing experience of the
be clean
operator
Tooth extraction
immediately after therapy
Microscopic evaluation of
residual calculus

Twenty-four
Group A: scaling and root 1.2% more residual calculus in
the explorer group
patients, a total of
planing plus explorer
70 molars
Group B: scaling and root Statistical significance only in
interproximal sites (pocket
planing plus Perioscopy
depth < 6 mm; 2.6%)
Treatment until root
No differences in residual
surface was considered
calculus in deep pockets, furcato be clean
Tooth extraction immedi- tion areas or on buccal lingual
surfaces
ately after therapy
Microscopic evaluation Treatment duration: endoscope
group showed a significant deof residual calculus
crease of time with increasing
experience of the operator

193

Meissner & Kocher

Table 2. (Continued)
Instrument

Reference

Design

Sample size

Perioscopy

(65)

Clinical and
histological
study

Six patients, a
total of 12 teeth

DetecTar

(23)

Randomized,
single-masked
study

Eight patients, a
total of 44 teeth
(176 surfaces)
Teeth extracted
immediately after
treatment
Microscopic
evaluation

Group A: no treatment,
calculus detection by
DetecTar
Group B: scaling and
root planing + DetecTar
until teeth were
considered to be clean
Control of the detection:
results after extraction

Group A: n = 96 surfaces;
79.4% sensitivity and 95.1%
specificity
Group B: n = 80 surfaces
(n = 58 initially positive,
n = 22 initially negative)

DetecTar

(24)

Randomized,
controlled
clinical study

One-hundred
patients with
plaque-associated
gingivitis

Group A (n = 50):
supragingival
debridement + oral
hygiene instruction
and motivation
Group B (n = 50):
supragingival
debridement + oral
hygiene instruction and
motivation + Detectar

Detectar group:
Plaque index (baseline 57.5%,
after 4 weeks 27.1%)
Bleeding on probing
(baseline 19.1%, after
4 weeks 7.1%)
Control group:
Plaque index (baseline 60.5%,
after 4 weeks 41.9%)
Bleeding on probing (baseline
23.1%, after 4 weeks 14.5%)

DetecTar

(32)

In vitro randomized study

Twenty extracted
periodontally
involved, calculuscovered teeth

Perioscan

(39)

In vitro study

Ten teeth, 200


measurements

Perioscan

(38)

In vitro study

Thirty-four teeth,
1363 measurements

Perioscan

(40)

In vitro study

Fifty extracted,
periodontally
involved, calculuscovered teeth

194

Method

Results

Histologically: formation
Scaling and root planing
of a long junctional epithelium,
plus Perioscopy
evidence of bone repair,
Tooth extraction
no signs of chronic
6 months after therapy
inflammation
Histologic evaluation
No control group

Specificity:
Teeth were scanned:
100% in blood
(a) with different
95-100% for all
working tip
angulations in saline
angulations of the
solution
fibreoptic (0, 10, 45
Sensitivity:
or 90)
Nearly 100% for all
(b) with different ambient
fluids (blood and saline angulations in saline solution
In blood:
solution)
100% for 90 angulation
Results were compared
89% for 45 angulation
with clinical and
70% for 10 to 0 angulation
histological findings
Detection results were Calculus and cementum were
compared with visual distinguishable with a sensitivity
findings on calculus of 88% and a specificity of 76%
and cementum
surfaces
Detection results were
compared with visual
findings, by moving the
instrument tip over the
calculus and cementum
surfaces

Calculus and cementum


were distinguishable with a
sensitivity of 76% and a
specificity of 86%

The smallest, recognizable


Calculus was removed
stepwise, in order to residual deposits had an average
determine the discrimi- diameter of 219 mm, an area of
21,600 mm2 and a circumfernative capability
ence of 748 mm; Sensitivity was
73% and specicity 80%

Calculus-detection technologies

Table 2. (Continued)
Instrument
Perioscan

Reference

Design

Sample size

Method

Results

(37)

In vivo
randomized,
clinical study

Sixty-three buccal
subgingival tooth
surfaces

Teeth were scanned


in situ
Detection results were
compared with visual
findings after extraction

Calculus and cementum


were distinguishable with a
sensitivity of 91% and a
specificity of 82%
The positive predictive
value was 0.59 and the
negative predictive value
was 0.97

ERL, Er:YAG laser; UI, ultrasonic instrumentation.

Fig. 2. Spectro-optical technology. The DetecTar (Dentsply Professional, York, PA, USA) uses a light-emitting
diode and fiberoptic technology to detect calculus.

Fig. 1. Endoscopy-based technology. The Perioscopy


(Perioscopy Inc., Oakland, CA, USA) uses a minimally
invasive miniature periodontal endoscope, which is
inserted into the periodontal pocket, to detect calculus.

small benefit only in interproximal sites, in particular


in single-rooted teeth with deep pockets, and in
multirooted teeth with relatively shallow pockets.

Spectro-optical technology
The spectro-optical approach to calculus detection
uses a light-emitting diode and fiberoptic technology,
and is currently used by only one device, the DetecTar (Dentsply Professional, York, PA, USA) (Fig. 2).
The characteristic spectral signature of subgingival

calculus, which is caused by absorption, reflection


and diffraction when irradiated by red light, is sensed
by an optical fiber and converted into an electrical
signal that is analyzed by a computer-processed
algorithm. The DetecTar device comes as a portable
cordless handpiece with a curved periodontal probe
that has millimeter markings to measure pocket
depths. Without any tactile pressure, the subgingival
root surface can be scanned by the instrument. As
soon as calculus is detected, the operator receives the
information on calculus localization by audible and
luminous signals.
Only a few investigations have evaluated spectrooptical technology as a diagnostic instrument in
periodontology. The ability to detect subgingival
calculus in vitro was tested in 20 freshly extracted
teeth affected by periodontitis, and the results were
compared with clinical and histological findings (32).
In addition, the influence of different working-tip
angulations (0, 10, 45 and 90) of the fiberoptic probe
and of different ambient fluids (blood and saline
solution) were studied. The specificity was only
slightly influenced by the type of irrigation fluid,
being 100% in blood and 95-100% in saline solution
for all angulations. The sensitivity in saline solution

195

Meissner & Kocher

was nearly 100% for all angulations. In blood, the


sensitivity decreased with smaller tip angulations
(100% sensitivity with angulation 90, 89% sensitivity
with angulation 45 and 70% sensitivity with angulation 100). The combination of saline solution as
the ambient fluid and a working-tip angulation of 90
which, however, cannot be achieved in the
periodontal pocket resulted in the most accurate
measurements.
A recent clinical study sought to determine the
utility of the spectro-optical technology for subgingival calculus removal (23). A total of 44 teeth (176
surfaces) were included in the study. In an untreated
control group, a total of 96 untreated surfaces were
scanned in vivo using the DetecTar. In the treatment
group, treatment was initiated upon obtaining positive signals from the spectro-optical device, and the
treatment was continued until no signal was elicited.
Clinical calculus findings were documented by visual
and microscopic examination after tooth extraction.
The control group showed a sensitivity of only 79.4%
and a specificity of 95.1%. Of 58 tooth surfaces that
initially showed calculus and which were consequently treated until they tested negative for calculus,
10 (17%) remained partly covered with calculus,
whereas 48 (83%) were completely calculus-free.
Nevertheless, nine (41%) of the 22 surfaces that were
initially identified as calculus-free (and therefore
untreated) did, in fact, harbor calculus. However, the
number of false-negative readings may have been
caused by incomplete surface scanning as a result of
limited access of the instrument and problems with
guiding the instrument. No sensitivity or specificity
data for the treatment group were calculated from
the published results. Additionally, the study only
recorded the clinical presence or absence of subgingival calculus deposits for each surface (without exact
localization on the respective surface), and a highly
heterogeneous group of surfaces, with pocket depths
ranging from 1 to 10 mm, was evaluated. Therefore,
false-negative results may have been caused by an
incomplete scanning process, technological limits of
the device, or a combination of both. These aspects
cannot be discriminated in vivo if the exact location of
the device during scanning is not definitively known.
Altogether, the utility of the spectro-optical
technology for calculus detection has not yet been
thoroughly investigated.

Autofluorescence-based technology
The ability of calculus to emit fluorescent light following irradiation with light of a certain wavelength

196

enables the detection of calculus, and several in vitro


studies have examined the autofluorescence of dental
root surfaces and calculus (8, 12, 18, 26, 30, 45). Oral
microorganisms and their metabolites (metal-free
porphyrins, metalloporphyrins and other chromatophores) are assumed to contain the fluorophores that
are emitted from dental calculus and from carious
lesions (14, 21, 29). Several distinct fluorescence
bands between 570 and 730 nm were identified on
calculus specimens, which could be elicited with light
of wavelength 400420 nm, but could not be found
on clean root surfaces (9). Another study found
characteristic autofluorescence emission peaks for
calculus and dentin caries at 700 and 720 nm,
respectively, which were elicited by light of wavelengths 635 and 655 nm, respectively (33). On
surfaces covered by bacterial cells or blood, the
autofluorescence intensity was reduced.
In order to differentiate calculus from the healthy
tooth surface, a fluorescence-ratio method based on
autofluorescence induced by a blue light-emitting
diode of 405 nm has been developed (48). Calculus
and healthy tooth surfaces exposed to light wavelengths of 487 and 628685 nm were used to create a
calculus parameter, R, which was selected to define a
relationship between the integrated intensities specific for calculus and for healthy teeth in the 628 to
685- and the 477 to 497-nm wavelength regions,
respectively. A cut-off threshold of R = 0.2 was able to
distinguish dental calculus from healthy teeth with
100% sensitivity and 100% specificity under various
experimental conditions in vitro.
A diagnostic instrument, based on different autofluorescence intensities after stimulation with red
light, claims to distinguish healthy from carious tooth
substance (Diagnodent; KaVo, Biberach, Germany)
(Fig. 3). An indium gallium arsenide phosphate
(InGaAsP)-based red laser diode (< 1 mW) sends light
with a wavelength of 655 nm through an optical fiber
onto the root surface, which is then induced to
fluoresce. The emitted fluorescent light returning
from the tooth tissue is captured by surrounding
optical fibers and transmitted to an integrated photo
diode, which serves as the fluorescence detector.
Optical effects caused by reflected light and ambient
light are eliminated by a band-pass filter and modulation of the fluorescent light, respectively. The
device was primarily developed for caries diagnosis
and launched as a stand-alone device about 10 years
ago. Based on a multitude of clinical studies, it is
considered to be a reliable caries detector on occlusal
and smooth surfaces, showing high levels of sensitivity (92.1%) and specificity (100%), a high level of

Calculus-detection technologies

Fig. 3. Autofluorescence-based technology. The DiagnodentTM Pen (KaVo, Biberach, Germany) is based on the
detection of different autofluorescence intensities after
stimulation with red light.

reproducibility (kappa value: in vitro, 0.9; in vivo,


0.9) and a good interexaminer and intra-examiner
agreement (21, 34, 35, 42, 46, 49, 58).
Later, the device was further refined to enable calculus detection. The fluorescence intensities are
measured, transformed and shown on a digital display as relative calculus-detection values from 0-99.
According to the manufacturer, values of 40 indicate
mineralized deposits, whereas values of between 5
and 40 indicate very small calcified plaque sites (not
further specified) or residual calculus following partial
cleaning, and values of 5 indicate a clean root surface. Values indicating calculus are indicated by a
beep with an increasing audiotone frequency as the
display value increases. The manufacturer thus provides a small-size device, which is claimed to be able
to detect both caries and calculus, and which can be
handled easily with no further training required.
The autofluorescence-based device for calculus
detection has been evaluated only in in vitro studies
so far, with any patient-derived clinical evidence
lacking. Surfaces of extracted periodontally involved
teeth, which were partly covered with calculus and
moistened with saline solution or blood, were scanned using the device (17, 31). The fluorescence
signals detected were compared with visual and
histological findings. The presence of calculus was
significantly correlated with a higher intensity of
fluorescence (17, 31). A median value of 6.2 was
obtained for clean root surfaces and a median value
of 57.7 was obtained for calculus, which was not
influenced by the presence of fluid. Additionally, high
reproducibility for measurements after 6 and 24 h
could be shown (31). The second study found relative

fluorescence values in air (cementum, 0.4; calculus,


54.1), in saline solution (cementum, 0.4; calculus,
60.7) and in blood (cementum, 2.1; calculus, 39.6).
With a cut-off value of 5, sensitivity and specificity in
all three media were 100% (17). Another study simulated a clinical situation based on a mannequin
model and compared the effectiveness of rootsurface instrumentation when supported by the
application of two different diagnostic instruments
(the autofluorescence-based system vs. a conventional explorer) (16). Forty extracted periodontally
involved teeth (120 surfaces for each diagnostic
group) were treated with conventional Gracey
curettes until this method indicated a clean root
surface. For multirooted teeth, calculus detection
using autofluorescence resulted in a significantly
smaller total area covered with residual calculus than
if diagnostics was based on a conventional explorer.
However, in single-rooted teeth, the two study groups
revealed a comparable amount of residual calculus.
In summary, when used in vitro, the autofluorescence-based system could differentiate between calculus and cementum with great reproducibility. In a
preclinical situation, a superior effect of the system
compared with manual use of an explorer could be
shown only on molars. The diagnostic value of the
autofluorescence-based system needs to be assessed
in the clinical setting, and its effect on treatment
outcomes determined.

Combined detection treatment


devices
Ultrasonic technology
Ultrasonic calculus-detection technology is based on
a conventional piezo-driven ultrasonic scaler and is
similar to the way that one might tap on the rim of a
glass with a spoon to identify cracks acoustically (28,
60). An insert at a conventional dental ultrasound
scaler receives short, weak impulses with a frequency
of about 50 Hz, which make the inserts distal tip
oscillate at a frequency that is dependent upon the
surface characteristics. The oscillations are conducted into the piezo-ceramic discs, which transform
the oscillations into voltage. The voltage level represents the intensity of the tip oscillation, while the
frequency stays the same. The overall signal, consisting of both the impulse stimulus and the impulse
response, is evaluated using a computerized system,
thereby generating information about a given surface
characteristic.

197

Meissner & Kocher

The ultrasonic device currently available (Perioscan; Sirona, Bensheim, Germany) (Fig. 4) provides
a detection mode to discriminate between calculus
deposits and clean roots, along with a treatment
mode that allows conventional ultrasonic treatment
at different power levels. When the ultrasonic tip
touches the tooth surface, the detection results are
indicated by a light signal integrated both in the
handpiece and in a display of the table unit (green
indicates cementum and blue indicates calculus).
When calculus is detected, an additional acoustic
signal sounds. The detection mode is only activated
when no scaling treatment is performed. The detection and treatment modes can be used successively
on the surface of the same tooth. If calculus deposits
are found, the root surface can be treated with a
higher power setting, whereas in the absence of calculus (thus requiring the systematic removal only of
biofilm), instrumentation can be performed at a
lower power setting. A prototype of the ultrasonic
device evaluated the calculus-detection capability
under laboratory conditions both in static tests
(yielding a sensitivity of 75% and a specificity of
82%) and during movements of the probing tip
(yielding a sensitivity of 88% and a specificity of
76%) (38, 39). The detection limit was further evaluated by gradually removing calculus from 50 extracted teeth until the system stopped discriminating
calculus deposits. Diameter, circumference and area
of the smallest recognizable deposit, and of the no
longer recognizable deposit, were measured, and a
cut-off point was determined. It could be demonstrated that calculus deposits with a diameter of
0.2 mm could still be recognized with a sensitivity of
73% and a specificity of 80% (40).

90% root
10% calculus

198

The only available study involving the clinical


application of this ultrasound tool tested the
accuracy by which calculus was detected (37).
In vivo calculus detection was determined on 63
subgingival surfaces and compared with visual
findings after tooth extraction. A prevalence of
calculus of 22.3% was found on the scanned surfaces, and calculus and cementum were discriminated with a sensitivity of 91% and a specificity of
82%. The positive and negative predictive values
were 0.59 and 0.97, respectively. The combined
application of the calculus-detection mode and
the ultrasonic removal of calculus remain to be
investigated.
To sum up, the combined detection-and-treatment
technology using ultrasound is a promising tool for
minimally invasive debridement (retaining cementum) and selective calculus removal, as shown by a
study employing an in vivo and ex vivo reconstruction technique. However, the long-term clinical outcome has not yet been investigated.

Laser-based technology
The benefit of laser application in nonsurgical periodontal therapy is still a matter of debate among
clinicians (4, 12, 51). Lately, out of a variety of other
types of lasers, the Er:YAG laser has been considered
to be the most promising for periodontal therapy (2,
3, 19). Its ability to ablate soft and hard tissue without
major thermal side effects qualifies the use of this
laser for periodontal therapy, and Er:YAG lasers at
different energy levels have been studied in various
in vitro and clinical trials. Er:YAG lasers are solidstate lasers that emit pulsed infrared light with a

Fig. 4. Ultrasound-based calculusdetection technology: Perioscan


(Sirona Dental Systems GmbH,
Bensheim, Germany). The principle
includes a fuzzy-logic-based detection mode employing ultrasound
feedback analysis and adds a treatment mode to the automated
calculus detection, which uses the
same tip.

Calculus-detection technologies

wavelength of 2940 nm, which is strongly absorbed


by virtually all biological tissues containing water.
The effect of Er:YAG lasers is based on photoablation.
The light-induced tissue evaporation results in water
release and a concomitant cooling effect on the surrounding tissue. However, when applied to dental
hard tissue, which contains a lower amount of water,
increased thermal effects can occur, and therefore
water irrigation is required (2).
The treatment effect of Er:YAG lasers (Keylaser 1 or
2; Kavo, Biberach, Germany) (Fig. 5) with regard to
calculus removal has been shown to be comparable
to conventional root debridement. No major thermal
damage was found if the laser was applied at lower
energy levels (radiation energy, 50160 mJ) and with
concomitant water irrigation (2, 15, 18, 19, 54). A
number of in vivo and in vitro studies have shown the
potential of Er:YAG lasers to create a biocompatible
root surface by removing the smear layer and lipopolysaccharides from the tooth surface, by promoting
the attachment of periodontal ligament fibroblasts
and by decreasing the bacterial load (1, 52, 66). By
contrast, studies have also reported increased tissue
removal, roughened surfaces and a lower yield of
calculus removal compared with hand instrumentation (3, 15, 18, 19). The effectiveness of calculus
removal seems to be dependent on the irradiation
energy level. However, the application of high energy
levels is also associated with increased and undesirable root-substance loss if applied to a healthy tooth
structure (2, 18, 19).
The only commercially available device (Keylaser3TM; KaVo) combines detection and treatment in a
feedback-controlled manner for selective removal of
calculus. The integrated calculus-detection device is
based on a 655-nm InGaAs diode laser for autofluorescence-based calculus detection (described above
as a stand-alone diagnostic tool), whereas a 2940-nm
Er:YAG laser is used for treatment. The Er:YAG laser is
only activated to emit light if a preselected autofluorescence threshold value for the diagnostic laser on
a scale of 099 is exceeded. As soon as the value falls
below the threshold, the Er:YAG laser turns off. This
combination of a diagnostic and a therapeutic laser
was designed to optimize calculus removal while
minimizing the undesired side effects of the Er:YAG
laser.
The feedback-controlled Er:YAG laser was recently
evaluated in in vitro and clinical studies to determine
how different fluorescence-classification thresholds
would influence the extent of calculus and cement
removal. Twenty teeth partly covered with calculus
and irrigated with water were treated from coronal to

Fig. 5. Laser-based combined detection and treatment


technology. The Keylaser 3 (KaVo, Biberach, Germany)
employs the same detection method depicted in Fig. 3, but
adds a treatment mode to it.

apical direction in contact irradiation mode with


pulsed infrared radiation [wavelength of 2.940 mm,
a chisel-shaped glass-fiber application tip (size
0.4 1.65 mm), 140 mJ per pulse, 10 Hz and calculated energy density of 17.2 mJ cm2) (30). The fluorescence threshold varied between 5 (recommended
by the manufacturer as the lowest threshold value)
and 1 in order to potentially increase sensitivity. Not

199

Meissner & Kocher

surprisingly, the amount of residual calculus depended on the laser fluorescence threshold levels. At
a threshold of 5, the median residual amount of calculus related to the baseline amount of calculus was
11% (minimum, 0%; maximum, 78%), whereas at a
threshold of 1, it was reduced to 0% (minimum, 0%;
maximum, 26%). However, the laser-treated residual
cementum was signicantly thinner (median, 80 lm)
than the untreated residual cementum (median, 90 lm;
P < 0.05). Thus, by reducing the threshold level to 1, the
sensitivity was increased at the expense of a reduced
specificity, as indicated by the increase of undesired
substance loss.
A different study compared the clinical and histological effects of conventional hand instrumentation
with fluorescence-controlled Er:YAG laser irradiation
at different device settings (55). Twenty-four periodontally involved single-rooted teeth were treated
in vivo and extracted after therapy. Laser treatment
consisted of fluorescence-controlled Er:YAG laser
irradiation under water irrigation (160 mJ per pulse,
chisel-shaped tip of 1.65 0.5 mm, calculated energy
density 19.4 J cm2 per pulse, 10 Hz). All mesial root
surfaces were treated in vivo under local anesthesia
until they were considered to be clean. After extraction,
the distal root surfaces were treated in vitro for comparison. Hand-instrumented teeth were treated accordingly. Clinically, the use of the Er:YAG laser in vivo
produced homogeneous and nearly smooth root surfaces
without visible traces of the tip. Histologically, calculus
had been selectively removed and no thermal damage
could be observed. The results were comparable to those
seen after the use of hand instruments. The treatments
with the Er:Yag laser and with the hand instruments
were found to be more effective in vitro than in vivo.
Laser treatment also resulted in the removal of an increased amount of cementum in vitro compared with
in vivo, whereas for hand instrumentation the in vitro
and in vivo results were comparable The reason for less
substance removal in vivo was assumed to be caused by
the restaining of the pocket tissue with blood and sulcus
fluid, which may have influenced the autofluorescence of
the dental hard tissue in vivo. However, by contrast,
different media (including blood and saline solution) did
not influence the autofluorescence intensity in vitro (17).
Another clinical study compared the clinical
benefit of autofluorescence-controlled Er:YAG laser
radiation with that of a special ultrasonic device
with vertical vibrations of the working tip (Vec rr, Bietigheim-Bissingen, Germany), and
torTM; Du
with hand instrumentation (53). Seventy-two singlerooted teeth that were scheduled for extraction from
12 patients were randomly treated by the laser (at one

200

of three energy levels: 100, 120 or 140 mJ per pulse,


10 Hz), the Vector ultrasound system, conventional
hand instruments, or remained untreated. Teeth
were instrumented in vivo under local anesthesia until
they were considered to be clean and then immediately
extracted for analysis. The ultrasound system left significantly smaller areas of residual calculus than the two
other therapies, but needed a significantly longer
instrumentation time than the laser and the hand
instruments. However, treatment with the feedbackcontrolled Er:YAG laser still resulted in significantly less
residual calculus and less root-surface alterations than
hand instrumentation.
A clinical study compared the microbiological
effects of the Er:YAG laser, hand instruments, sonic
scalers and ultrasonic scalers (13). The controlled,
randomized, single-blinded clinical trial included 72
periodontal patients who had at least one site per
quadrant with a pocket depth of > 4 mm, bleeding
on probing and bone loss of at least 33%. The four
quadrants per patient were randomly assigned to one
of the following four debridement modalities: hand
instruments, a feedback-controlled Er:YAG laser
(Keylaser3; 160 mJ per pulse, 10 Hz, water irrigation,
chisel-shaped tips of 0.5 1.65 and 0.5 1.1 mm), a
sonic scaler (SONICflexs system LUX 2003 L; KaVo)
or a piezoelectric ultrasonic scaler (Piezon Master
400; EMS, Nyon, Switzerland). Subgingival plaque
samples were obtained at baseline and at 3 and
6 months postoperatively. All four treatments resulted in a significant reduction in the amounts of
Porphyromonas gingivalis, Prevotella intermedia,
Tannerella forsythia and Treponema denticola after
3 months. Laser and sonic instrumentation failed to
significantly reduce the amount of Aggregatibacter
actinomycetemcomitans. Six months post-treatment,
the amount of test bacteria had increased in all study
groups.
Another set of clinical trials compared the clinical
outcome of periodontal treatment by a feedbackcontrolled Er:YAG laser or ultrasonic instrumentation (56). Single-rooted and multirooted teeth with
pocket depths of > 4 mm were randomly treated in
a split-mouth design either by a feedback-controlled
Er:YAG laser (160 mJ per pulse, 10 Hz, chisel-shaped
tip of 1.65 0.5 mm, calculated energy density
136 mJ per pulse; or chisel-shaped tip of 1.1
0.5 mm, calculated energy density 114 mJ per pulse)
or by an ultrasonic device (Cavitron Select; Dentsply, Konstanz, Germany) (56). At baseline, and 3 and
6 months post-treatment, plaque index, bleeding on
probing, pocket depth, gingival recession and clinical attachment level were measured at six sites per

Calculus-detection technologies

tooth. Deep pockets showed a tendency to experience more gingival recession, to gain more clinical
attachment level and to retain more residual pocket
depth compared with moderately deep pockets.
Bleeding on probing and clinical attachment level
improved significantly in both treatment groups
after 6 months compared with baseline. However,
statistically significant differences could not be
observed between the two types of treatment, suggesting that treatment with the Er:YAG laser was
comparable with, but probably not superior to,
ultrasonic instrumentation (56). This conclusion is
in agreement with a subsequent clinical study that
compared the microbiological and short-term clinical effects after Er:YAG laser debridement vs. ultrasonic treatment (62). Twenty patients with at least
two pockets with a depth of > 5 mm in each jaw
were included in the study. The pockets were randomized to receive either feedback-controlled
Er:YAG laser treatment (160 mJ per pulse, 10 Hz,
chisel-shaped tip of 1.1 0.5 mm, water irrigation)
or piezoelectric ultrasonic treatment (Piezon Master
400; EMS). Clinical attachment level gain and pocket
depth reduction after 1 month were significantly
higher in the laser group (mean pocket depth
reduction, 0.9 mm; mean clinical attachment level
gain, 0.5 mm) than in the ultrasonic group [mean
pocket depth reduction, 0.5 mm (P < 0.05); mean
clinical attachment level gain, 0.06 mm (P < 0.01)],
whereas 4 months after retreatment, no significant
differences were detected between the two treatment modalities (mean pocket depth reduction:
laser, 1.1 mm; ultrasonic, 1.0 mm; and mean clinical
attachment level gain: laser, 0.6 mm; ultrasonic,
0.4 mm). Both treatment modalities yielded a similar reduction of the subgingival microflora after
4 months.
In conclusion, clinical and histological studies have
shown that laser-based detection and treatment of
calculus can effectively remove subgingival calculus
and preserve root substance. However, the results
were comparable with hand and ultrasonic debridement, and controlled long-term clinical studies are
lacking.

Summary
A number of different technologies have been
incorporated into dental devices for the purpose of
identifying and selectively removing dental calculus.
Some of these new approaches for calculus removal
show promising results under optimum in vitro

conditions. Histological and microscopic findings


after in vivo use point to the potential for some of
these technologies to support or replace conventional
subgingival scaling. Published studies evaluating
clinical parameters, however, exist only for the
ultrasound- and laser-based devices, which combine
calculus detection and treatment. Moreover, controlled randomized clinical trials are lacking for all
currently commercially available dental devices that
are used to identify and selectively remove dental
calculus.
All studies starting out with teeth treated in vivo
and then investigated after extraction have the same
problem in common, namely that clinical parameters
such as pocket depth, gingival recession and clinical
attachment level are assumed to be associated with a
comparable prevalence of calculus. This might not
always be the case and therefore a bias of uncertain
magnitude is introduced, especially if different studies and methods are compared. Moreover, it is
questionable whether the claimed improvement in
calculus detection in fact has resulted in selective
calculus removal and a concomitant preservation of
cementum. Without histologic examination, it is
impossible to decide whether cementum has actually
also been removed (50). In the case of the laser-based
detection and treatment device, for instance, histological analysis unveiled that the thorough removal of
calculus also resulted in an unwanted increase in the
amount of cementum removed.
A common problem of the stand-alone diagnostic
devices is that the application of these instruments
requires the systematic scanning of the entire subgingival tooth surface, and, in the case of positive
calculus detection, the detected calculus has to be
located using the therapeutic scaling instrument.
Identifying the exact location of the calculus may be
difficult, thus potentially leading to over-treatment or
under-treatment. This problem relates to the skills of
the clinician rather than to features of the instrument. The combined detection and treatment
instruments aim to overcome this problem.
The influence of operator skills on the outcome
variable has been shown previously and should
always be considered when evaluating the utility of a
particular method of scaling (8). Two different scenarios are conceivable: an experienced and trained
clinician will manage more easily the application of
advanced diagnostic procedures, such as the endoscopy-based system, and thus obtain better results
than an inexperienced operator. Alternatively, a clinician who is highly experienced in traditional scaling methods may achieve less additional benefit by

201

Meissner & Kocher

using supportive detection devices than a beginner or


a modestly skilled clinician, who may overcome a
lack of manual dexterity by using a supportive diagnostic system. These aspects have not been addressed in the published literature.
The fiberoptic detection technology shows potential to be a helpful tool in periodontal therapy, but
needs to be studied in clinical studies in direct
comparison with established scaling techniques. The
fiberoptic device currently available is somewhat
difficult to handle and requires additional time and
skills of the operator, especially when used simultaneously with scaling and root planing.
Data on the clinical utility of a spectro-optical
device for scaling and root planing are scarce.
Promising results were shown regarding the sensitivity and specificity of calculus detection in vitro.
Whether or not a spectro-optical device is useful for
calculus detection needs to be evaluated in a clinical
setting. To our knowledge, a spectro-optical device is
not currently available for dental use.
To date, published data on autofluorescence-based
detection technology are only available from in vitro
and mannequin model studies. The autofluorescence-based system was found to be superior to
scaling and root planing alone only for multirooted
teeth, possibly because of their complicated root
configuration, which makes conventional diagnostics
more difficult. Calculus removal in single-rooted
teeth yielded similar results with and without the use
of the autofluorescence-based system. Its effectiveness in clinical situations and its impact on clinical
parameters remains to be investigated.
It may be easy for clinicians to learn how to use
and apply the ultrasonic-driven combined detection
and treatment device because it is similar to the
familiar scaling technique. To provide reliable data
on the benefits of the device, clinical studies are
necessary to investigate changes in pocket depth,
clinical attachment level, bleeding on probing and
occurrence of hypersensitivity after treatment compared with conventional methods.
The laser-based calculus-detection and treatment
technology has shown promising results with respect
to histology and certain clinical parameters in one
study, which, however, was limited to single-rooted
teeth. As the total number of cases in the published
literature is still small, additional studies are necessary to evaluate the clinical benefit of this technology.
Taken together, despite promising laboratory
research results, the new technology-assisted
periodontal treatments have yet to show clinical
superiority in comparison with conventional scaling.

202

Clinical studies are necessary to assess if the use of


these devices can improve long-term treatment outcome, with consequences of smaller residual probing
depth, a reduced need for periodontal surgery and
less hypersensitivity after treatment.

Acknowledgment
The work on Perioscan was supported by grants from
r Bildung und Forschung
the Bundesministerium fu
(BMBF 01 EZ 0025, BMBF 01 EZ 0026) and
from Sirona, Bensheim, Germany. T. Kocher and
G. Meissner have served as consultants to Sirona.

References
1. Ando Y, Aoki A, Watanabe H, Ishikawa I. Bactericidal effect
of erbium YAG laser on periodontopathic bacteria. Lasers
Surg Med 1996: 19: 190200.
2. Aoki A, Ando Y, Watanabe H, Ishikawa I. In vitro studies on
laser scaling of subgingival calculus with an erbium:YAG
laser. J Periodontol 1994: 65: 10971106.
3. Aoki A, Miura M, Akiyama F, Nakagawa N, Tanaka J, Oda S,
Watanabe H, Ishikawa I. In vitro evaluation of Er:YAG laser
scaling of subgingival calculus in comparison with ultrasonic scaling. J Periodontal Res 2000: 35: 266277.
4. Aoki A, Sasaki KM, Watanabe H, Ishikawa I. Lasers in
nonsurgical periodontal therapy. Periodontol 2000 2004:
36: 5997.
5. Avradopoulos V, Wilder RS, Chichester S, Offenbacher S.
Clinical and inflammatory evaluation of Perioscopy on
patients with chronic periodontitis. J Dent Hyg 2004: 78:
3038.
6. Biller IR, Kerber PE. Reliability of scaling error detection.
J Dent Educ 1980: 44: 206210.
7. Blomlof L, Lindskog S, Appelgren R, Jonsson B, Weintraub
A, Hammarstrom L. New attachment in monkeys with
experimental periodontitis with and without removal of
the cementum. J Clin Periodontol 1987: 14: 136143.
8. Brayer WK, Mellonig JT, Dunlap RM, Marinak KW, Carson
RE. Scaling and root planing effectiveness: the effect of root
surface access and operator experience. J Periodontol 1989:
60: 6772.
9. Buchalla W, Lennon AM, Attin T. Fluorescence spectroscopy of dental calculus. J Periodontal Res 2004: 39: 327332.
10. Cobb CM. Non-surgical pocket therapy: mechanical. Ann
Periodontol 1996: 1: 443490.
11. Cobb CM. Clinical significance of non-surgical periodontal
therapy: an evidence-based perspective of scaling and root
planing. J Clin Periodontol 2002: 29(Suppl. 2): 616.
12. Cobb CM. Lasers in periodontics: a review of the literature.
J Periodontol 2006: 77: 545564.
13. Derdilopoulou FV, Nonhoff J, Neumann K, Kielbassa AM.
Microbiological findings after periodontal therapy using
curettes, Er:YAG laser, sonic, and ultrasonic scalers. J Clin
Periodontol 2007: 34: 588598.

Calculus-detection technologies
14. Dolowy WC, Brandes ML, Gouterman M, Parker JD, Lind J.
Fluorescence of dental calculus from cats, dogs, and
humans and of bacteria cultured from dental calculus. J Vet
Dent 1995: 12: 105109.
15. Eberhard J, Ehlers H, Falk W, Acil Y, Albers HK, Jepsen S.
Efficacy of subgingival calculus removal with Er:YAG laser
compared to mechanical debridement: an in situ study.
J Clin Periodontol 2003: 30: 511518.
16. Folwaczny M, Heym R, Mehl A, Hickel R. The effectiveness
of InGaAsP diode laser radiation to detect subgingival
calculus as compared to an explorer. J Periodontol 2004: 75:
744749.
17. Folwaczny M, Heym R, Mehl A, Hickel R. Subgingival calculus detection with fluorescence induced by 655 nm InGaAsP diode laser radiation. J Periodontol 2002: 73: 597601.
18. Folwaczny M, Mehl A, Haffner C, Benz C, Hickel R. Root
substance removal with Er:YAG laser radiation at different
parameters using a new delivery system. J Periodontol
2000: 71: 147155.
19. Frentzen M, Braun A, Aniol D. Er:YAG laser scaling of
diseased root surfaces. J Periodontol 2002: 73: 524530.
20. Geisinger ML, Mealey BL, Schoolfield J, Mellonig JT. The
effectiveness of subgingival scaling and root planing: an
evaluation of therapy with and without the use of the
periodontal endoscope. J Periodontol 2007: 78: 2228.
21. Hibst R, Paulus R, Lussi A. Detection of occlusal caries by
laser fluorescence: basic and clinical investigation. Med
Laser Appl 2001: 16: 205213.
22. Hughes FJ, Smales FC. Immunohistochemical investigation
of the presence and distribution of cementum-associated
lipopolysaccharides in periodontal disease. J Periodontal
Res 1986: 21: 660667.
23. Kasaj A, Moschos I, Rohrig B, Willershausen B. The effectiveness of a novel optical probe in subgingival calculus
detection. Int J Dent Hyg 2008: 6: 143147.
berpru
24. Kasaj A, Sculean A, Willershausen B. Studie zur u
fung der wirksamkeit des detectar-systems hinsichtlich
einer verbesserung der compliance und der mundhygiene.
Parodontologie 2004: 15: 343353.
25. Kepic TJ, OLeary TJ, Kafrawy AH. Total calculus removal:
an attainable objective? J Periodontol 1990: 61: 1620.
26. Kocher T, Konig J, Hansen P, Ruhling A. Subgingival polishing compared to scaling with steel curettes: a clinical
pilot study. J Clin Periodontol 2001: 28: 194199.
27. Kocher T, Ruhling A, Momsen H, Plagmann HC. Effectiveness of subgingival instrumentation with power-driven
instruments in the hands of experienced and inexperienced operators. A study on manikins. J Clin Periodontol
1997: 24: 498504.
28. Kocher T, Strackeljan J, Behr D. Feasibility of computerassisted recognition of different dental hard tissues. J Dent
Res 2000: 79: 829834.
29. Konig K, Flemming G, Hibst R. Laser-induced autofluorescence spectroscopy of dental caries. Cell Mol Biol
(Noisy-le-grand) 1998: 44: 12931300.
30. Krause F, Braun A, Brede O, Eberhard J, Frentzen M, Jepsen
S. Evaluation of selective calculus removal by a fluorescence feedback-controlled Er:YAG laser in vitro. J Clin
Periodontol 2007: 34: 6671.
31. Krause F, Braun A, Frentzen M. The possibility of detecting
subgingival calculus by laser-fluorescence in vitro. Lasers
Med Sci 2003: 18: 3235.

32. Krause F, Braun A, Jepsen S, Frentzen M. Detection of


subgingival calculus with a novel LED-based optical probe.
J Periodontol 2005: 76: 12021206.
33. Kurihara E, Koseki T, Gohara K, Nishihara T, Ansai T,
Takehara T. Detection of subgingival calculus and dentine
caries by laser fluorescence. J Periodontal Res 2004: 39: 59
65.
34. Lussi A, Hibst R, Paulus R. DIAGNOdent: an optical method
for caries detection. J Dent Res 2004. 83 Spec No C: C8083.
35. Lussi A, Megert B, Longbottom C, Reich E, Francescut P.
Clinical performance of a laser fluorescence device for
detection of occlusal caries lesions. Eur J Oral Sci 2001:
109: 1419.
36. Mandel ID, Gaffar A. Calculus revisited. A review. J Clin
Periodontol 1986: 13: 249257.
37. Meissner G, Oehme B, Strackeljan J, Kocher T. Clinical
subgingival calculus detection with a smart ultrasonic device: a pilot study. J Clin Periodontol 2008: 35: 126132.
38. Meissner G, Oehme B, Strackeljan J, Kocher T. In vitro
calculus detection with a moved smart ultrasonic device.
J Clin Periodontol 2006: 33: 130134.
39. Meissner G, Oehme B, Strackeljan J, Kocher T. Influence of
handling-relevant factors on the behaviour of a novel
calculus-detection device. J Clin Periodontol 2005: 32: 323
328.
40. Meissner G, Oehme B, Strackeljan J, Kocher T. A new system to detect residual subgingival calculus: in vitro detection limits. J Clin Periodontol 2006: 33: 195199.
41. Michaud RM, Schoolfield J, Mellonig JT, Mealey BL. The
efficacy of subgingival calculus removal with endoscopyaided scaling and root planing: a study on multirooted
teeth. J Periodontol 2007: 78: 22382245.
42. Nemes J, Csillag M, Toth Z, Fazekas A. Reproducibility of
the laser fluorescence method for the diagnosis of occlusal
caries. Clinical study. Fogorv Sz 2001: 94: 3336.
43. Nyman S, Sarhed G, Ericsson I, Gottlow J, Karring T. Role of
diseased root cementum in healing following treatment
of periodontal disease. An experimental study in the dog.
J Periodontal Res 1986: 21: 496503.
44. Nyman S, Westfelt E, Sarhed G, Karring T. Role of diseased root cementum in healing following treatment of
periodontal disease. A clinical study. J Clin Periodontol
1988: 15: 464468.
45. Pihlstrom BL, Ammons WF. Treatment of gingivitis and
periodontitis. Research, Science and Therapy Committee
of the American Academy of Periodontology. J Periodontol
1997: 68: 12461253.
46. Pinelli C, Campos Serra M, de Castro Monteiro Loffredo L.
Validity and reproducibility of a laser fluorescence system
for detecting the activity of white-spot lesions on free
smooth surfaces in vivo. Caries Res 2002: 36: 1924.
47. Pippin DJ, Feil P. Interrater agreement on subgingival
calculus detection following scaling. J Dent Educ 1992: 56:
322326.
48. Qin YL, Luan XL, Bi LJ, Lu Z, Sheng YQ, Somesfalean G,
Zhou CN, Zhang ZG. Real-time detection of dental calculus
by blue-LED-induced fluorescence spectroscopy. J Photochem Photobiol B 2007: 87: 8894.
49. Ross G. Caries diagnosis with the DIAGNOdent laser: a
users product evaluation. Ont Dent 1999: 76: 2124.
50. Ruhling A, Wulf J, Schwahn C, Kocher T. Surface wear on
cervical restorations and adjacent enamel and root

203

Meissner & Kocher

51.

52.

53.

54.

55.

56.

57.

cementum caused by simulated long-term maintenance


therapy. J Clin Periodontol 2004: 31: 293298.
Schwarz F, Aoki A, Becker J, Sculean A. Laser application in
non-surgical periodontal therapy: a systematic review.
J Clin Periodontol 2008: 35: 2944.
Schwarz F, Aoki A, Sculean A, Georg T, Scherbaum W,
Becker J. In vivo effects of an Er:YAG laser, an ultrasonic
system and scaling and root planing on the biocompatibility of periodontally diseased root surfaces in cultures of
human PDL fibroblasts. Lasers Surg Med 2003: 33: 140
147.
Schwarz F, Bieling K, Venghaus S, Sculean A, Jepsen S,
Becker J. Influence of fluorescence-controlled Er:YAG laser
radiation, the Vector system and hand instruments on
periodontally diseased root surfaces in vivo. J Clin Periodontol 2006: 33: 200208.
Schwarz F, Sculean A, Berakdar M, Georg T, Reich E, Becker
J. Clinical evaluation of an Er:YAG laser combined with
scaling and root planing for non-surgical periodontal
treatment. A controlled, prospective clinical study. J Clin
Periodontol 2003: 30: 2634.
Schwarz F, Sculean A, Berakdar M, Szathmari L, Georg T,
Becker J. In vivo and in vitro effects of an Er:YAG laser, a
GaAlAs diode laser, and scaling and root planing on periodontally diseased root surfaces: a comparative histologic
study. Lasers Surg Med 2003: 32: 359366.
Sculean A, Schwarz F, Berakdar M, Romanos GE, Arweiler
NB, Becker J. Periodontal treatment with an Er:YAG laser
compared to ultrasonic instrumentation: a pilot study.
J Periodontol 2004: 75: 966973.
Sherman PR, Hutchens LH Jr, Jewson LG, Moriarty JM,
Greco GW, McFall WT Jr. The effectiveness of subgingival
scaling and root planning. I. Clinical detection of residual
calculus. J Periodontol 1990: 61: 38.

204

58. Shi XQ, Welander U, Angmar-Mansson B. Occlusal caries


detection with KaVo DIAGNOdent and radiography: an in
vitro comparison. Caries Res 2000: 34: 151158.
59. Stambaugh RV, Myers G, Ebling W, Beckman B, Stambaugh
K. Endoscopic visualization of the submarginal gingiva
dental sulcus and tooth root surfaces. J Periodontol 2002:
73: 374382.
60. Strackeljan J, Behr D, Kocher T. Fuzzy-pattern recognition
for automatic detection of different teeth substances. Fuzzy
Sets and Systems 1997: 85: 275286.
61. Tan B, Gillam DG, Mordan NJ, Galgut PN. A preliminary
investigation into the ultrastructure of dental calculus and
associated bacteria. J Clin Periodontol 2004: 31: 364369.
62. Tomasi C, Schander K, Dahlen G, Wennstrom JL.
Short-term clinical and microbiologic effects of pocket
debridement with an Er:YAG laser during periodontal
maintenance. J Periodontol 2006: 77: 111118.
63. Van der Weijden GA, Timmerman MF. A systematic review
on the clinical efficacy of subgingival debridement in the
treatment of chronic periodontitis. J Clin Periodontol 2002:
29(Suppl 3): 5571.
64. White DJ. Dental calculus: recent insights into occurrence,
formation, prevention, removal and oral health effects of
supragingival and subgingival deposits. Eur J Oral Sci 1997:
105: 508522.
65. Wilson TG Jr, Carnio J, Schenk R, Myers G. Absence of
histologic signs of chronic inflammation following closed
subgingival scaling and root planing using the dental
endoscope: human biopsies - a pilot study. J Periodontol
2008: 79: 3641.
66. Yamaguchi H, Kobayashi K, Osada R, Sakuraba E, Nomura
T, Arai T, Nakamura J. Effects of irradiation of an erbium:YAG laser on root surfaces. J Periodontol 1997: 68: 1151
1155.

Copyright of Periodontology 2000 is the property of Wiley-Blackwell and its content may not be copied or
emailed to multiple sites or posted to a listserv without the copyright holder's express written permission.
However, users may print, download, or email articles for individual use.

Anda mungkin juga menyukai