Anda di halaman 1dari 22

Downloaded from http://rspa.royalsocietypublishing.

org/ on September 26, 2018

Proc. R. Soc. A
doi:10.1098/rspa.2011.0601
Published online

Toner-print removal from paper by long and


ultrashort pulsed lasers
BY DAVID RICARDO LEAL-AYALA1 , J. M. ALLWOOD1, *,
M. SCHMIDT2 AND I. ALEXEEV2
1 Low
Carbon Materials Processing Group, Department of Engineering,
University of Cambridge, Trumpington Street, Cambridge CB2 1PZ, UK
2 Lehrstuhl für Photonische Technologien, Paul-Gordan-Straße 3,
91052 Erlangen, Germany

Toner-print removal from paper would allow paper to be re-used instead of being recycled,
incinerated or disposed of in landfill. This could significantly reduce the environmental
impact of paper production and use. Previous work on the subject has explored the
applicability of ultraviolet, visible and infrared (IR) lasers under nanosecond pulses for
toner removal. This article expands on this work by testing a wider range of ultrafast
and long-pulsed lasers. Results from 10 distinct laser set-ups are used to propose an
operating window for the toner-removal process. Colour analysis under the L∗ a ∗ b ∗ colour
space, scanning electron microscope examination and attenuated total reflectance–Fourier
transform IR spectroscopy measurements of the outcome show that, with the right laser,
it is possible to remove toner from paper to enable its re-use. Theoretical models to
predict the laser ablation of toner are discussed, and, while imperfect, provide sufficient
evidence to support a physical explanation of toner ablation.
Keywords: paper recycling; toner print; laser ablation; material efficiency;
reduction of paper demand

1. Introduction: why is paper re-use relevant?

In a highly energy-efficient industry, such as the pulp and paper sector, there
is little room for improvements that would yield a considerable reduction of its
energy consumption and associated impact on climate change. Reducing paper
demand appears unlikely as paper consumption has increased steadily throughout
recent decades and this trend is expected to continue in the near future, as
suggested by Hekkert et al. (2002). Employing carbon neutral fuels and increasing
the energy efficiency of manufacturing processes are widespread policies in the
pulp and paper industry, as indicated by the Confederation of European Paper
Industries (CEPI 2009). Material recovery through recycling eliminates the
forestry step from the life cycle of paper and eradicates emissions arising from
*Author for correspondence (jma42@cam.ac.uk).
Electronic supplementary material is available at http://dx.doi.org/10.1098/rspa.2011.0601 or via
http://rspa.royalsocietypublishing.org.

Received 20 October 2011


Accepted 16 February 2012 1 This journal is © 2012 The Royal Society
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

2 D. R. Leal-Ayala et al.

incineration or landfill dumping. However, there is modest room for improvement


if it is considered that the current European recycling rate of 66 per cent is close
to the theoretical maximum rate of 81 per cent estimated by CEPI (2006). What
would happen if paper was unprinted and re-used instead of being recycled?
This action would remove four steps from the life cycle: forestry, pulping, paper-
making and disposal by incineration or landfill. Counsell & Allwood (2007) have
estimated that this could translate into a reduction of up to 95 per cent CO2 -
equivalent (CO2 -eq) emissions per tonne of office paper, in comparison with only
76 per cent achieved by recycling.
This article discusses the applicability of an alternative paper re-use method
that employs laser radiation to remove toner print from paper. Previous work on
the subject by Leal-Ayala et al. (2011) demonstrated that there is a particular
combination of laser parameters that can remove certain toner-print types from
paper using visible light at 532 nm with long pulses of 4 ns. It is known from
laser–matter theory that laser ablation of solid materials follows different physical
mechanisms depending on the duration of the laser pulses. Long pulses in the
nanosecond or slower time scales first heat up the surface of the target material
to its melting point and then to its vaporization temperature. In contrast, during
ablation of solid surfaces exposed to ultrafast pulses in the pico- and femtosecond
scales, material removal occurs by direct vaporization away from the surface
(quick solid–vapour transition with only a brief liquid phase). In this case, thermal
conduction into the target can be neglected because of the very short time scales
involved. Therefore, smaller pulse lengths can, in theory, diminish the negative
thermal effects generated by long pulses on the substrate. The objective of this
article is to answer the following questions:

— Can toner print be effectively removed from paper by laser ablation in


pulse regimes other than those explored in previous work? And, if so, can
these methods produce paper sheets of high enough quality to replace new
paper?
— What is the operating window of the laser removal process for the complete
range of lasers tested?
— What happens to toner and paper when they interact with ultrafast laser
pulses? How is this process different from toner and paper interaction with
long-pulsed lasers?
— Can this process be modelled and controlled?

This study starts by reviewing previous work on the subject and explaining
the relevant laser–matter interaction theory. The experimental and analytical
methodology is summarized in §3. Section 4 presents a series of experimental trials
aimed at identifying the most suitable type of laser for toner-print removal from a
range of ultrafast and long-pulsed lasers. Analyses of the laser–paper interaction
and removal mechanisms are provided in §5 together with a definition of the
operating window of the laser process and two ablation depth predictive models.
The certainty about the results is discussed in §6 together with the economical
and environmental implications of laser removal and possibilities for future work.

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 3

2. Background: previous work on toner removal by laser ablation

The idea of using lasers to remove toner print from paper, although not new,
has not been fully developed until now. This is evidenced by a survey performed
by Counsell & Allwood (2006), who stated that most of the previous research in
this area exists only in the form of patents that provide no details about their
success. The closest source of literature on the effects that laser radiation might
have on cellulose-based paper comes from the artwork restoration field. A detailed
review of the literature on this subject has been performed by Leal-Ayala et al.
(2011), suggesting that ultraviolet (UV) laser radiation has a degrading effect
over both blank and soiled paper, decolouring cellulose fibres through photo-
oxidative reactions. In addition, infrared (IR) light is also considered harmful
because of the extended damage and discoloration that it produces on paper as
a result of photothermal reactions originating from heat transfer from the soil
into the substrate. The effects of both UV and IR radiation on paper seem to
contrast with those of visible light, which has been successfully used to clean
ancient manuscripts and documents.
In theory, paper could be cleaned and re-used without damage to its substrate
if a laser fluence level can be found below the ablation and destruction threshold
of the paper and above that required to ablate toner. A study by Ihori et al. (2009)
has explored this idea by testing long-pulsed IR and visible lasers working at 1064
and 532 nm, respectively, on printed paper samples. They determined that, while
the IR laser burned the paper and reduced its moisture, visible radiation with
energy fluence of 1.2 J cm−2 was capable of removing toner while maintaining
paper whiteness to a level that would allow it to be re-printed. The exact damage
and degree of re-usability of paper after treatment is not explicit in their results.
It seems that the whiteness level achieved after toner removal was not the same
as that of blank un-lased paper.
In the study by Leal-Ayala et al. (2011), five different UV, visible and IR lasers
with nanosecond pulses were applied to a range of toner–paper combinations
to determine their ability to remove toner. They found that it is possible to
remove toner from paper without considerable substrate damage when using
visible radiation at 532 nm with 4 ns pulses and energy fluences under 1.6 J cm−2 .
UV and IR lasers created higher damage to cellulosic paper. The authors
suggested the possibility that UV and IR laser radiation act as a catalyst that
accelerates paper degradation mechanisms (photochemical and photothermal) as
a result of the absorption of high-energy photons and the generation of localized
heating. Regarding the ablation mechanism, they considered that the high-energy
absorption of the toner results in a localized temperature increase that leads
to the thermal degradation of its polymer fraction (for toner samples roughly
composed of 50 wt% polyester resin and 50 wt% iron oxide), detaching the rest
of its components from paper. This led to the conclusion that toner ablation
under long laser pulses at a wavelength of 532 nm is a photothermally dominated
process. Based on this assumption, the authors suggested that their results might
be improved if the negative thermal effects exerted by laser radiation on paper
could be avoided.
As explained by Liu et al. (1997), absorption of laser energy by the target
material can be achieved by linear or nonlinear absorption mechanisms. The
former includes direct ionization of valence electrons with an ionization potential

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

4 D. R. Leal-Ayala et al.

lower than photon energy while the latter refers to multi-photon and avalanche
ionization processes. In the case of opaque materials such as toner, linear
absorption dominates under long-pulsed ablation with low intensities while
nonlinear absorption becomes more important during high-intensity ultrashort
pulse lengths. In both situations, the material is heated as a result of Joule
heating. The amount of heat generated is directly related to the pulse duration.
Longer pulses lead to a stronger energy transfer and a higher heat generation.
During long pulses, there is enough time for the thermal wave to propagate
into the material and create a relatively large melted layer while only a small
portion of the material is evaporated. Shorter pulse lengths generate higher
laser peak powers, increasing the electrons’ peak temperatures to the range
of a few to tens of electrovolts (1 eV = 11 600◦ K), as explained by Liu et al.
(1997), while ion and material lattice temperatures remain relatively low. Owing
to this thermal gradient and the short duration of the pulses, electron–ion
energy transfer continues after the laser pulse is over, heating the ions to
much higher temperatures than do long pulses, vaporizing a large portion of
the material subjected to laser radiation. The interaction time is so short that
heat has little time to diffuse into the material, reducing the energy losses and
the heat-affected volume. The melt layer remains small as most of the heated
material is vaporized. These characteristics of ultrashort laser pulses could be
beneficial during the ablation of toner from paper and need to be explored
given that earlier work on this subject and related paper conservation studies
have only dealt with lasers working in the long-pulsed nanosecond regime. In
addition, little is known about the physical mechanisms behind toner removal
by laser ablation at distinct pulse regimes. The following sections address these
knowledge gaps.

3. Methodology

A total of 10 laser set-ups were tested in this study: six in the ultrashort-
pulsed regime and four in the long-pulsed category. Ultrafast ablation tests were
performed on a Fuego Integrated Picosecond MOPA laser (Time-Bandwidth
Products) and a Spitfire Pro XP Ultrafast Amplifier (Spectra-Physics). The
former runs at a constant pulse width of 10 ps with adjustable wavelengths
between 366, 532 and 1064 nm. The Spitfire laser device works at wavelengths of
266, 400 and 800 nm with variable pulse lengths ranging from 20 fs to 3 ps. Long-
pulsed tests were performed on a QuikLaze 50ST2 laser (New Wave Research) for
tests with 4 ns pulses at 532 nm, an SPI Photonics System for trials at 1064 nm
with 40 ns pulses, and a Navigator II YHP40 OEM Laser System for tests at 532
and 355 nm with adjustable pulse lengths between 29 and 75 ns.
Tests were performed on uniform rectangular areas printed with HP LaserJet
Q1338A-AC-D black toner on white, uncoated, wood-free (lignin-free) 80 g m−2
Canon copy paper. The exact composition of the paper and fillers is unknown
as that information is not commonly released by Canon. It is assumed that a
typical sheet of office paper is formed by 70–90% cellulose fibres, while the rest
is a combination of fillers (typically clays), whiteners (typically titanium oxide)
and additives related to strength and water-absorption properties. The toner

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 5

composition reported by HP (2004) consists of 40–50 wt% polyester resin, 40–


50 wt% ferrous iron oxide and 1 wt% amorphous silica. During each test, the
samples were placed on the laser bed in the focal plane of the beam and moved in
a raster pattern to generate a rectangular scan across the printed sample surface.
The full range of parameters tested in this study for femtosecond, picosecond and
nanosecond lasers are contained in the electronic supplementary material. Only
one variable was changed at a time while the rest were set at a default value. The
variables included wavelength, energy fluence, pulse frequency, pulse width, scan
speed, spot area and number of horizontal passes. The best parameters identified
for each laser are shown in §4.
All test samples were analysed under an optical microscope in order to select
the best results. These were qualitatively evaluated under a scanning electron
microscope (SEM) and subjected to a quantitative colour analysis obtained by
scanning them with a CanoScan Lide 25 scanner from Canon and converting
the resulting images to the LAB colour space using IMAGEJ software. The L∗ ,
a ∗ and b ∗ colour variables from the LAB colour model were measured with
the same software. Under this system, L∗ defines the lightness of the colour
(L∗ = 0 for black and L∗ = 100 for white), a ∗ measures green and magenta tones
(negative = green, positive = magenta) and b ∗ quantifies the yellow and cyan
colours in a sample (negative = cyan, positive = yellow). Every scanned sample
was accompanied by two colour references (black and white) which were used to
calibrate the process. L∗ a ∗ b ∗ variables from each colour reference in each sample
were measured and compared with the previously determined values to ensure
that the scanner repeatability remained at an acceptable level. The scanned
sample would be accepted if there were no significant variations in the reference
values. Ten measurements were taken from each lased paper sample under study
(L∗ , a ∗ and b ∗ for each variable). The values reported in this study are average
values with their corresponding standard deviation.
Further analysis was performed through Fourier transform IR spectroscopy
with attenuated total reflectance (FTIR–ATR). This technique can be used
to evaluate whether exposure to laser radiation produces any changes in the
atomic structure of paper—and hence any damage—by measuring the changes
that occur in a reflected IR beam when the beam comes into contact with a
paper sample. ATR–FTIR spectra were collected at a 1 cm−1 resolution between
525 and 4000 cm−1 using a Bruker Optics FTIR Tensor 27 coupled with a Pike
Miracle ATR single reflection accessory, fitted with a diamond crystal and using
an angle of incidence of 45◦ . A high-pressure clamp was used to ensure good
contact between the diamond (active area of 1.8 mm) and the paper samples
in order to obtain high-quality spectra. All paper samples were conditioned in
accordance to the T402 sp-08 TAPPI standard (standard conditioning and testing
atmospheres for paper, board, pulp hand sheets and related products). Pre-
conditioning consisted of introducing the samples in a forced ventilation oven
at a temperature of 39◦ C for a 2 h period. The oven was operated in a room at
55 per cent relative humidity (RH) and 23◦ C. Under these conditions, room air
drawn into the oven resulted in an RH between 20 and 30 per cent inside the
oven (according to the information given in the T402 sp-08 standard). After pre-
conditioning, the samples were exposed to the conditioning atmosphere (55% RH
and 23◦ C) for 4 h in order for them to come into equilibrium with the atmosphere.
After this, the samples were transported as quickly as possible to the testing

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

6 D. R. Leal-Ayala et al.

(a) (b) (c) (d) (e) (f)

Figure 1. Ultrashort-pulsed laser samples—(a) 800 nm at 1000 fs, (b) 400 nm at 500 fs, (c) 266 nm at
120 fs, (d) 1064 nm at 10 ps, (e) 532 nm at 10 ps, and (f ) 355 nm at 10 ps. (Online version in colour.)

atmosphere where the ATR–FTIR machine was located. The testing atmosphere
was measured as 54 per cent RH and 21◦ C. This procedure ensured that all
samples possessed similar humidity levels during testing.

4. Toner removal by ultrafast lasers

This section presents the best results obtained from toner-print removal
experiments using UV, visible and IR lasers under pico- and femtosecond pulse
regimes. Additional results with long-pulsed lasers in the nanosecond regime are
shown in §4b. Further evaluation of the results is presented in §4c.

(a) Results
Real size test samples and microscopic images taken from them are shown in
figures 1 and 2. Colour measurements from all lasers under study are also included
in figure 2 while the optimum experimental parameters used for each laser are
summarized in tables 1 and 2. The results suggest that visible and IR light with
10 picosecond pulses are the most successful in removing toner while minimizing
paper damage. The rest of the lasers evaluated inflicted considerable damage on
cellulosic fibres.
It can be seen from figure 2 that, although UV, visible and IR lasers (266,
400 and 800 nm, respectively) under femtosecond pulses fully remove toner print
from paper, they also inflict considerable damage to the paper substrate. All
three lasers left linear marks on cellulose fibres which suggest that a layer of the
material was removed. UV radiation at 355 nm with 10 ps pulses does not cut
cellulose fibres as the femtosecond lasers did but it produced a dramatic change
in their physical appearance without achieving full toner removal.
Better results were obtained with the visible and IR picosecond lasers. Figures 1
and 2 show that IR radiation at 1064 nm with 10 ps pulses successfully removed
toner from the sample while the values of all three colour variables remained
very close to those of un-lased paper (only a minor decrease of 5 points in L∗
and increments of 3 and 2 for a ∗ and b ∗ , respectively). Despite this, a comparison
of the 1064 nm sample at 10 ps with blank un-lased paper suggests that cellulose
fibres lost some of their bonding and appear less densely packed after exposure
to the laser treatment. This signifies that laser radiation at 1064 nm has slightly
damaged the paper areas under the toner print.

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 7

L=100 ± 0, a=0 ± 0, b=0 ± 0 L=84 ± 6, a=2 ± 1, b=0 ± 1 L=84 ± 5, a=2 ± 3, b=1 ± 2 L=90 ± 4, a=1 ± 1, b=6 ± 2
(a) (i) (ii) (iii) (iv)

200 mm 200 mm 200 mm 200 mm

(b) (i) (ii) (iii) (iv)

20 mm 20 mm 20 mm 20 mm

L=95 ± 2, a=3 ± 1, b=2 ± 1 L=95 ± 1, a=2 ± 1, b=6 ± 1 L=83 ± 6, a=1 ± 1, b=0 ± 1


(a) (v) (vi) (vii)

200 mm 200 mm 200 mm

(b) (v) (vi) (vii)

20 mm 20 mm 20 mm

Figure 2. (a) Optical microscope (×20) and (b) SEM images from ultrashort-pulsed trials—(i)
blank paper, (ii) 800 nm at 1000 fs, (iii) 400 nm at 500 fs, (iv) 266 nm at 120 fs, (v) 1064 nm at
10 ps, (vi) 532 nm at 10 ps, and (vii) 355 nm at 10 ps. (Online version in colour.)

Table 1. Optimum parameters for femtosecond laser tests.

femtosecond tests

wavelength (nm) 800 400 266


pulse width (fs) 1000 500 120
fluence (J cm−2 ) 1.5 × 10−2 4.2 × 10−2 9.6 × 10−3
peak power (W) 3 × 106 1.5 × 107 2.5 × 107
pulse frequency (kHz) 10 10 10
scan speed (mm s−1 ) 139 279 18
line spacing (mm) 150 150 150
number of passes 6 5 2
spot area (m2 ) 2 × 10−8 1.8 × 10−8 3.1 × 10−8

Visible radiation at 532 nm and 10 ps pulses is also capable of achieving full


toner removal, while it appears to create no mechanical damage in cellulose fibres
(figure 2). Colour analysis reveals that there is little difference between the L∗
and a ∗ values from this sample and white un-lased paper. However, there is an

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

8 D. R. Leal-Ayala et al.

L=79 ± 3, a=11 ± 1, b=28 ± 3 L=95 ± 3, a=1 ± 1, b=1 ± 1 L=80 ± 4, a=2 ± 1, b=2 ± 1 L=77 ± 5, a=4 ± 1, b=2 ± 1
(a) (i) (ii) (iii) (iv)

200 mm 200 mm 200 mm 200 mm

(b) (i) (ii) (iii) (iv)

20 mm 20 mm 100 mm 20 mm

Figure 3. (a) Optical microscope (×20) and (b) SEM images from long-pulsed trials in the
nanosecond regime—(i) 1064 nm at 40 ns, (ii) 532 nm at 4 ns, (iii) 532 nm at 29 ns, and (iv) 355 nm
at 75 ns. A different scale has been used on (b (iii)) in order to present a better view of the damage.
(Online version in colour.)

Table 2. Optimum parameters for picosecond laser tests.

picosecond tests

wavelength (nm) 1064 532 355


pulse width (ps) 10 10 10
fluence (J cm−2 ) 0.2 1 0.08
peak power (W) 2.5 × 106 5.6 × 106 2.5 × 106
pulse frequency (kHz) 0.4 2.2 0.4
scan speed (mm s−1 ) 18 28 18
line spacing (mm) 150 150 150
number of passes 2 2 2
spot area (m2 ) 1.1 × 10−8 5.7 × 10−9 3.1 × 10−8

increment of 6 units in b ∗ , which means that there is minor discoloration or


yellowing of the paper after exposure to laser radiation. This yellowing can be
appreciated from a naked-eye visual examination of the sample.

(b) Toner removal with long-pulsed lasers


Microscopic images from the best long-pulsed laser removal results are shown
in figure 3. The samples shown in figures 3a(i,ii),b(i,ii) were obtained with
the experimental parameters previously reported by Leal-Ayala et al. (2011).
Although the four lasers shown in figure 3 were able to remove toner print
from paper, they all interact in distinct ways with the paper substrate. Sample
3a(i),b(i) suffered significant yellowing after toner ablation with a 1064 nm laser
under 40 ns pulses. It can be seen from its SEM image that the fibres lost some
of their bonding when compared with blank un-lased paper. The same effect can
be appreciated in sample 3a(iii),b(iii) exposed to visible radiation at 532 nm with

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 9

Table 3. Optimum experimental parameters for the 532 nm laser with 4 ns pulses.

pulse width (ns) 4


fluence (J cm−2 ) 1.6
peak power (kW) 92.5
pulse frequency (Hz) 15
scan speed (mm s−1 ) 150
line spacing (mm) 50
number of passes 1
spot area (m2 ) 2.25 × 10−8

29 ns pulses. A very different outcome was obtained with UV light at 355 nm


and 75 ns pulses in sample 3a(iv),b(iv). This type of laser damaged the paper
fibres considerably, producing a complete change in their morphology. The best
result was obtained with visible radiation at 532 nm with 4 ns pulses, shown in
sample 3a(ii),b(ii). This laser left no residual toner on the ablated area and no
appreciable damage on cellulose fibres. The exact parameters used for this test
are summarized in table 3.

(c) Evaluation
The first question established at the beginning of this study can be answered
at this point: toner print can be effectively removed from paper by laser ablation
in pulse regimes distinct from those previously explored by Leal-Ayala et al.
(2011). In particular, visible and IR radiation at 532 and 1064 nm with 10 ps
pulses are outstandingly successful in removing toner from paper. The second
question—‘Can these methods produce paper sheets of high enough quality to
replace new paper?’—deserves careful consideration. Most of the tested lasers
generated sufficient physical damage on cellulose fibres to be considered as
unsuitable solutions for the problem under study. One of the best results was
obtained with visible radiation at 532 nm and 4 ns pulses, in agreement with
previous work. Two options stand out from the rest of the tested lasers: IR and
visible light at 10 ps. The former leaves the paper with high enough quality to
be re-used in terms of appearance, but the mechanical damage that it inflicts
on cellulose fibres means that the number of times it could be re-used will be
limited by its mechanical resistance to the treatment. On the contrary, visible
light at 10 ps leaves cellulose fibres in a good structural shape but it generates
a minor whiteness loss. In this case, paper can be re-used but the amount of
times this process could be repeated would be limited by its physical appearance.
In summary, a conditional positive answer can be given to the second question
stated here the best results obtained under visible and IR laser radiation with
10 ps pulses produce paper sheets of high enough quality to replace new paper,
as concluded from colorimetric and SEM analyses, suggesting that it might be
technically feasible to achieve paper re-use after ultrafast laser ablation of toner
print. However, the number of times that paper could be re-used will be limited by
the detrimental side-effects of the removal treatment. This needs to be explored
in further work.

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

10 D. R. Leal-Ayala et al.

bleaching under
print and blank
areas
IR
fibre bonding/ yellowing
cohesion loss under
print
sliced
fibres
wavelength

vis
yellowing
under print no damage
and blank
areas

UV morphology
transformation

1e–13 1e–12 1e–11 1e–10 1e–9 1e–8


pulse length (s)

Figure 4. Relationship between wavelength, pulse length and paper damage (areas are
approximate). Dark grey circles, tested laser set-up.

5. Defining an operating window: laser–matter interaction and paper damage

Figure 4 and table 4 classify all laser set-ups tested in the previous section based
on the relation between their wavelength type, pulse length and the damage that
they exert on paper. A detailed discussion of this classification is presented in
§5a–c.
The threshold fluence for ablation of a certain material will become lower if the
pulses are shortened as this will translate into much higher peak powers. This is
confirmed by the power and fluence trends in figure 5, which show a more specific
view on how the optimum ablation fluence and peak power vary as a function
of the pulse length for all lasers tested in the visible spectrum. A clear trend
can be appreciated: smaller pulses lead to smaller optimum fluences and higher
peak powers. Similar graphs of fluence versus peak power for UV and IR tests
are contained in the electronic supplementary material.
Although it would be extremely difficult to test all possible combinations
of wavelength, pulse width, fluence and peak power, the information shown
in figure 4 is a representative sample of a very wide laser range covering UV,
visible and IR radiation in long and ultrashort pulse regimes. Therefore, the
data in this figure can be considered as the operating window of the toner-
removal process and represent the answer to the second question stated at the
beginning of this study. An examination of this operating window suggests that
the tested lasers produced varied effects during their interaction with toner
and paper. This leads to the third question set in §1: what happens to toner

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 11

Table 4. Effects of wavelength and pulse length on paper and their main degradation reactions.
Bold text indicates the optimum laser set-up found in this investigation.

wavelength pulse length physical effects main degradation


(nm) (s) on paper reactions

UV 266 120 × 10−15 sliced fibres photothermal degradation


paper morphology leading to localized
transformation vaporization
direct photolysis by energetic
photons.
355 10 × 10−12 paper morphology photochemical degradation
transformation leading to ionization events
and bond breakage.
355 75 × 10−9 paper morphology photochemical degradation:
transformation photo-oxidation
visible 400 500 × 10−15 sliced fibres photothermal degradation
leading to localized
vaporization
532 10 × 10−12 yellowing under print and photochemical and
blank areas photothermal degradation
leading to chromophore
generation
532 4 × 10−9 no damage: optimum none
regime
532 29 × 10−9 fibre bonding/cohesion loss photothermal degradation
leading to ablation of
secondary materials
IR 800 1000 × 10−15 sliced fibres photothermal degradation
leading to localized
vaporization
1064 10 × 10−12 fibre bonding/cohesion loss photochemical and
bleaching under print and photothermal degradation
blank areas leading to chromophore
generation
ablation of secondary
materials
1064 40 × 10−9 fibre bonding/cohesion loss thermal degradation through
yellowing under print heat diffusion from toner

and paper when they interact with ultrafast laser pulses? How is this process
different from their interaction with long-pulsed lasers? This section addresses
these queries.

(a) Paper damage under ultraviolet radiation and femtosecond ablation


mechanisms
UV radiation has a strong negative effect over cellulose under any
experimental conditions. This behaviour could be attributed to the creation
of photochemical degradation reactions in cellulose as a result of UV light

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

12 D. R. Leal-Ayala et al.

optimum experimental parameters (visible)


Ppeak fluence
16 6
14
5
12
peak power (MW)

fluence (J cm–2)
4
10
8 3
6
2
4
1
2
0 0
1e–13 1e–11 1e–09 1e–07
pulse length (s)

Figure 5. Peak power and optimum fluence versus pulse length for visible results.

absorption. Kolar et al. (2000) measured the absorption of UV and visible light
by cellulose and found that, although most of the light in the visible band is
scattered, absorption is greatly increased in the UV region. Photons in this region
are more energetic than those on the visible and IR fractions of the light spectrum.
Direct photolysis (chemical process by which molecules are broken down into
smaller units through the absorption of light) of an atomic covalent bond can
occur if the energy of an absorbed photon is higher than that of the bond. Typical
bonds in cellulose include C−C, C−O and C−H links. The approximate energies
for these bonds are 347, 359 and 414 kJ mol−1 according to Kolar et al. (2000).
These energies correspond to UV light wavelengths of 347, 333 and 289 nm,
respectively. Given that the shorter the wavelength, the more energetic its photons
become, the photon energy at 266 nm used in this study is higher than the energies
from the three types of covalent bonds mentioned before. Absorption of this kind
of photons could result in direct cellulose degradation by photolysis. In the case
of laser tests at 355 nm, this wavelength has an energy level slightly lower than
that of C−C, C−O and C−H bonds and photolysis seems unlikely. It is also
possible that absorption of UV photons at 266 and 355 nm results in excitation
of electrons in a chemical bond leading to ionization events and thus facilitating
photochemical degradation reactions, as mentioned by Kolar et al. (2000). Both
situations result in considerable deterioration of cellulose fibres.
A similar suggestion has been provided by Kaminska et al. (2004) when
exploring the use of a 355 nm laser with 6 ns pulses to remove a mixture of charcoal
and dust contamination from hand-made cellulose paper. They concluded that
UV radiation produced photochemical reactions in cellulose that led to severe
cellulose degradation by photo-oxidation.
The level of impairment in cellulose was more severe for the 266 nm laser with
120 fs pulses, as evidenced from SEM analysis in figure 2. As explained by Liu
et al. (1997), shorter pulse lengths generate higher electron peak temperatures,
which allows for electron–ion energy transfer to continue after the laser pulse
is over, heating the ions to much higher temperatures than long laser pulses

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 13

1.0

0.9
transmittance (normalized)

0.8
OH hydrogen bond
dissociation
0.7
3344
C–O ether bond
dissociation
0.6 1028

0.5
4000

3500

3000

2500

2000

1500

1000

500
cm–1

Figure 6. ATR–FTIR spectra from UV laser samples (normalized with respect to the most intense
band in the spectrum near the 4000 cm−1 limit). Black solid line, normal paper; dashed black line,
10 ps at 355 nm; grey solid line, 120 fs at 266 nm.

and vaporizing a large portion of the material subjected to laser radiation.


This effect was repeated for all wavelengths tested in the femtosecond regime
(266 nm under 120 fs pulses, 400 nm under 500 fs pulses and 800 nm under 1000 fs
pulses), where the elevated peak power magnitudes led to highly localized
vaporization, leaving no traces of any heat-affected zone on the remaining
cellulose fibres.
Analysis of the samples shown in figures 2 and 3 under ATR–FTIR indicates
that there is a considerable reduction in hydrogen bonds from OH groups in
cellulose after treatment with UV lasers. This is depicted in figure 6 by the
increase in the transmittance (T) at 3344 cm−1 , which signifies that hydrogen
bonds from OH groups responsible for attaching cellulose molecular chains
together are dissociated during the laser ablation process, hence altering the
normal structure of cellulose. The femtosecond laser at 266 nm resulted in
the highest OH hydrogen bond loss followed by the 355 nm with 10 ps pulses
and 355 nm with 75 ns pulses. This suggests that there is a highly localized
photochemical degradation effect induced by UV light on paper, in addition
to the photochemical reactions discussed before, and also suggests that shorter
pulse lengths (and therefore higher peak powers) result in higher photochemical
degradation of cellulose. In addition, there is a considerable difference in the band
at 1028 cm−1 , which corresponds to C−O ether bonds, suggesting a dissociation
of bonds as a result of additional photochemical degradation reactions. These
factors make UV laser unsuitable for toner removal.

(b) Visible spectrum: from paper yellowing to optimum removal


Even if cellulose scatters most laser light, it is still possible that other
components of paper absorb a certain amount of radiation and allow for chemical

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

14 D. R. Leal-Ayala et al.

1.0

0.9
transmittance (normalized)

formation of
oxidized groups
1724
0.8
OH hydrogen bond
dissociation
0.7
3344
C–O ether
bond
0.6
dissociation
1028
0.5
4000

3500

3000

2500

2000

1500

1000

500
cm–1

Figure 7. ATR–FTIR spectra from visible laser samples (normalized with respect to the most
intense band in the spectrum near the 4000 cm−1 limit). Black solid line, normal paper; dashed
black line, 10 ps at 532 nm; dashed grey line, 500 fs at 400 nm; grey solid line, 4 ns at 532 nm.

changes to take place, as suggested by Kolar et al. (2000). Typical office paper
contains elements such as hemicelluloses, lignin, fillers, pigments, dyes and sizing
agents in addition to cellulose fibres. These materials are also sensitive to laser
radiation and can initiate photo-induced reactions in the substrate, leading to
photodegradation. The SEM image shown in figure 3 corresponding to a sample
ablated under 532 nm with 29 ns pulses shows that cellulose fibres lose their
cohesion and appear less densely packed than blank un-lased paper. In other
words, the secondary components of paper (fillers, pigments, dyes and so on)
have disappeared from the sample. The 29 ns pulses appear long enough for
secondary materials to absorb sufficient laser energy to allow photothermal
reactions to occur, resulting in ablation of these materials while leaving cellulose
untouched owing to its distinct absorption rate and ablation threshold. The
same effect is no longer appreciated when the pulse length is reduced to 4 ns
into the optimum ablation region shown in figure 4. Reducing the pulse length
from 29 to 4 ns results in a slight peak power increase and a considerable toner
ablation threshold fluence reduction, as shown in figure 5, suggesting that a power
threshold has to be met in order to avoid the negative effects observed under
29 ns pulses while an upper pulse length limit restricts the overall acceptable
energy fluence.
Figure 7 indicates that light at 532 nm with 10 ps has a strong interaction
with paper. This sample suffered a considerable loss of hydrogen bonds from OH
groups responsible for attaching cellulose molecular chains together, represented
by the increment in the transmittance at the 3344 cm−1 band. It also suffered
the greatest increment in transmittance at the 1028 cm−1 band corresponding
to C−O ether bonds, suggesting the existence of photochemical degradation
reactions that result in the dissociation of both hydrogen bonds from OH groups

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 15

and C−O bonds. These reactions could be responsible for the yellowing effect
observed on both printed and blank paper after exposure to this kind of radiation.
As explained by Liu et al. (1997), the light absorption mechanism changes
towards nonlinear absorption at ultrashort pulse lengths with high intensities.
This means that mechanisms such as multi-photon absorption and avalanche
ionization become more dominant. It is likely that these absorption mechanisms
result in photolysis or excitation of electrons in chemical bonds which could
facilitate photochemical reactions. In addition, absorption of 532 nm light with
10 ps leads to high electron peak temperatures owing to the high peak power
employed. Therefore, both photochemical and photothermal reactions could be
induced on cellulose.
It is known (Daniels 1996) that paper degrades naturally through three main
reactions: acid-catalysed hydrolysis of cellulose molecules, oxidative degradation
induced from atmospheric oxygen and light, and thermal degradation that leads
to chemical bond breakage as the temperature is increased. Artificial paper
ageing tests have demonstrated that these reactions become more intense if
temperature is incremented. Carter (1989) has suggested that natural paper
degradation reactions are accompanied by the generation of chromophores
(discoloured components) that are responsible for the natural yellowing that
occurs on paper under ordinary degradation. Photochemical and photothermal
degradation reactions resulting from laser–paper interaction are similar to
the natural deterioration mechanisms enlisted by Daniels (1996), so it seems
plausible that laser radiation works as a catalyst that accelerates paper
degradation as a result of multi-photon absorption and excitation of electrons
in chemical bonds (leading to bond dissociation), and the generation of
localized heating (as denoted in figure 7), in a way similar to accelerated
ageing tests. This could in turn quicken the generation of yellow chromophores
and consequently be the reason behind the yellowing observed on paper after
laser radiation at 532 nm with 10 ps. All these make this laser unsuitable for
toner-print removal.

(c) Infrared radiation: paper discoloration and fibre-bonding loss


Long-pulsed IR light at 1064 nm with 40 ns pulses removed toner print from
paper at the cost of damaging and yellowing the paper areas under the print, as
seen in figure 3. This effect was not observed over blank areas of the paper also
subjected to laser radiation. This leads to the idea that during long IR pulses
there is enough time for heat to diffuse from the toner layer exposed to laser
radiation onto the paper substrate. The transferred heat works as a catalyst for
thermally induced cellulose degradation reactions accompanied by the production
of yellow chromophores. A similar effect has been reported by Kolar et al. (2002)
when using a 1064 nm laser to remove charcoal powder from purified cotton
cellulose. The substrate under the soil became yellowed after laser treatment. This
discoloration was attributed to the creation of yellow chromophores in cellulose
as a consequence of photothermal degradation driven by heat transfer from the
carbon soil to cellulose.
The yellowing under the print is no longer observed when toner is ablated under
the same wavelength with shorter pulses of 10 ps, although both cases produce
similar mechanical damage on cellulose fibres as can be appreciated from figures 2

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

16 D. R. Leal-Ayala et al.

1.0
transmittance (normalized)

0.9

0.8
OH hydrogen bond
dissociation
0.7
3344
C–O ether bond
dissociation
0.6
1028

0.5
4000

3500

3000

2500

2000

1500

1000

500
cm–1

Figure 8. ATR–FTIR spectra from the best laser toner-removal samples (normalized with respect
to the most intense band in the spectrum near the 4000 cm−1 limit). Black solid line, normal paper;
dashed black line, 10 ps at 532 nm; dashed grey line, 10 ps at 1064 nm; grey solid line, 4 ns at 532 nm.

and 3. IR light at 1064 nm with 10 ps pulses has a bleaching/whitening effect


over paper areas under the print and blank paper. The short interaction time
and elevated peak power magnitude lead to a much quicker ablation of the toner
layer, leaving no time for heat diffusion between toner and paper. This eliminates
the yellowing under the print observed with longer laser pulses. The bleaching
of the paper can be attributed to the same phenomenon observed during tests
with the 532 nm and 10 ps laser, where nonlinear absorption mechanisms such as
multi-photon absorption and avalanche ionization become present. ATR–FTIR
spectra of the three best results obtained in this study are shown in figure 8.
It can be seen that the spectra from both lasers at 1064 and 532 nm with 10 ps
pulses are quite similar at the 3344 and 1028 cm−1 bands, suggesting that both
have a strong and comparable interaction with the paper substrate. Therefore,
the formation of discoloured chromophores during laser ablation at 1064 nm with
10 ps seems likely, as was the case for laser radiation at 532 nm with 10 ps pulses,
leading to bleaching of the substrate.
The third question established in §1 of this study can be answered at this point:

— The high peak powers employed during all femtosecond tests led to
elevated substrate temperatures that resulted in strong mechanical damage
as a consequence of highly localized vaporization of material.
— UV radiation is not suitable for toner-print removal mostly owing to the
creation of photochemical deterioration reactions in cellulose which lead
to direct cellulose degradation by photolysis or by excitation of electrons
in chemical bonds leading to ionization events.
— The difference between ablation under green light at 532 nm with 4 and
29 ns pulses can be attributed to the fact that the duration of the latter

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 17

is long enough for secondary materials in paper to become ablated while


a reduction of pulse length to 4 ns leads to a higher peak power and lower
energy fluence combination which prevents any elements in paper from
being ablated. The same wavelength under shorter pulses of 10 ps achieves
toner removal but discolours the paper substrate as a result of accelerated
chromophore generation, in a way similar to accelerated paper ageing tests.
— The same effect is true for IR radiation under 10 ps pulses. In contrast,
an IR laser at 1064 nm with 40 ns pulses damages and yellows paper as
a result of heat diffusion from toner to the substrate during the ablation
process, facilitating the thermal degradation of cellulose.

In addition, the following topics could be further explored in future work:

— It is necessary to explore whether colour changes observed in the paper


substrate after laser radiation could be related to damage of the optical
brighteners commonly used in office paper. Future work could include UV–
visible fluorescence spectra measurements of the lased samples to verify
this.
— The changes seen at the surface of some samples could be related to
thermal degradation of CaCO3 or TiO2 minerals commonly used as paper
fillers. The fact that the paper is uncoated reduces the potential presence
of these minerals to the filler content only. The low percentage of fillers
in the paper would suggest that most of the appreciable changes seen
at the surface are due to cellulose-related reactions. Despite this, X-ray
fluorescence analysis could be used to verify the presence of these minerals
in future work.

(d) Ablation depth prediction


This section examines the applicability of two theoretical models to predict the
toner ablation depth obtained using the three best lasers identified in this study.

(i) Ablation under a 532 nm laser with 4 ns and 10 ps pulses


Ablation depth prediction for this particular wavelength under 4 ns pulses
was explored by Leal-Ayala et al. (2011), who applied a one-dimensional model
developed by Ready (1965) to predict the toner ablation rate as a function of
energy fluence. Ready’s model is based on the assumption that the laser energy
absorbed at the surface of an object quickly heats up the material to its melting
point and then to the vaporization temperature. At this point, the energy per
unit mass deposited in the material becomes larger than the specific heat of
evaporation and the laser starts supplying the latent heat of vaporization to a
thin layer of the surface, creating a vapour front that is removed immediately.
The surface, initially at z = 0, moves inward with time, reaching a position z(t)
at time t. A schematic view of this process and a detailed theoretical explanation
can be found in the electronic supplementary material. Ready’s adapted model
is described by
x·(A·H )·tp
D= , (5.1)
r(C ·Tv + Lv )

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

18 D. R. Leal-Ayala et al.

1.8
1.6
1.4
1.2
depth (mm)

1.0
0.8
0.6
0.4
0.2

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6


fluence (J cm–2)

Figure 9. Theoretical and measured values of ablation depth versus fluence for a 532 nm laser with
4 ns and 10 ps pulses. Solid line, model prediction; diamonds, measured values under 10 ps pulses;
squares, measured values under 4 ns pulses.

where x represents the effective energy fraction, A is the absorption fraction,


tp equals pulse time in seconds, heat flux H is measured as W m−2 , r stands
for material density (kg m−3 ), C is the specific heat capacity (J(kg◦ K)−1 ), Tv is
the vaporization temperature (◦ K) and Lv stands for latent heat of vaporization
(J kg−1 ). This model can predict the ablation depths obtained when varying the
energy fluence of a 532 nm laser with 4 ns and 10 ps pulses used to remove HP
LaserJet Q1338A-AC-D black toner from paper. The experimental values used
by Leal-Ayala et al. (2011) and the ablation depth predictions are shown in
table 5 and figure 9, respectively. The power H and the pulse width tp were
modified according to the characteristics of the lasers under study while the rest
of the parameters remained constant. Considering that the toner under study
is mainly composed of 50 wt% polyester resin and 50 wt% iron oxide, a key
assumption made is that full toner removal can be achieved by degrading the
polyester fraction, as this action would detach the rest of the toner elements from
the paper. Assuming a 50 per cent heat flow loss to the substrate (rule of thumb),
the remaining energy is distributed between the two material fractions that
compose the toner. The polyester resin fraction degrades at a lower temperature
than iron oxide, which means that only 25 per cent of the energy is available to
detach the toner.

(ii) Ablation under a 1064 nm laser with 10 ps pulses


The model proposed by Ready (1965) fails to predict the ablation rate
obtained with a 1064 nm laser working with 10 ps pulses. This could be related
to the fundamental difference in the way visible and IR light are absorbed and
inter-relate with toner at an atomic level. Although there is no general consensus

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 19

Table 5. Ablation depth theoretical model variables and input values (532 nm laser with 4 ns and
10 ps pulses).

x 0.25
A 0.95
H (W m−2 ) variable
tp (s) variable
P (kg m−3 ) 2200
C (J(kg◦ K)−1 ) 618
Tv (◦ K) 673
Lv (J kg−1 ) 4.81 × 105

in the literature regarding IR ablation depth models for polymer composites


such as toner, the logarithmic dependence of the ablation depth on the laser
pulse fluence, based on the Beer–Lambert absorption law, is well known for the
ablation of polymers and has also been demonstrated for ultrafast ablation of
metal targets. This is exemplified by the work of Schmidt et al. (2002) and
Preuss et al. (1995), respectively. Toner is neither a metal nor a pure polymer,
but a polymer composite containing a high fraction of electrically conductive iron
oxide. The ablation rate described by the Beer–Lambert law is valid as long as the
optical penetration depth exceeds the thermal penetration depth, as this means
that the optical effects would govern the ablation process. The use of picosecond
pulses allows this method to be employed as very low thermal penetration depths
are achieved. The model is described as
1 F
D= ln , (5.2)
a Fth

where D stands for ablation depth, a is the absorption coefficient, F is the applied
fluence and Fth is the threshold fluence for ablation. The last has been determined
experimentally as 0.035 J cm−2 . The absorption coefficient a can be obtained by
simple curve fitting using equation (5.2) and the depth measurements obtained
experimentally as shown in figure 10. The value of the absorption coefficient a
has been determined as 4.4 × 106 cm−1 , giving an optical penetration depth of
0.0022
√ mm. The thermal penetration depth can be estimated from the formula
2 (Dt), where D refers to thermal diffusivity (estimated at 1.1 × 10−3 cm2 s−1 for
toner) and t is the pulse length (10 ps), giving a thermal penetration of 0.0021 mm.
Since the optical penetration depth is practically the same as this value, it can
be said that the Beer–Lambert absorption law is valid to describe the process.
The results shown in figures 9 and 10 demonstrate that the ablation depth
is directly determined by the energy fluence employed, regardless of the pulse
width value. In any case, the pulse length remains a critical parameter owing
to the distinct impacts that it has on the paper substrate under the toner, as
discussed before. The data presented in this section confirm that it is possible to
control the laser ablation of toner and provide an answer to the fourth question
set in §1 of this study. The use of these models for other types of lasers and toners
is an area that remains open and needs to be explored in future work.

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

20 D. R. Leal-Ayala et al.

1.0
0.9
0.8
0.7
depth (mm)

0.6
0.5
0.4
0.3
0.2
0.1
0
0.01 Fth 0.1 1 10
fluence (J cm–2)

Figure 10. Theoretical and experimental values of ablation depth versus fluence for a 1064 nm laser
with 10 ps pulses. Diamonds, measured values under 10 ps pulses; solid line, model prediction.

6. Certainty about best results and future work

Additional research by Leal-Ayala (2012) has looked at the commercial feasibility


of laser un-printing by exploring the mechanical damage and ageing stability of
paper after laser treatment, and the safety, energy and economic performance of
the process. They found that the best laser identified in this study (4 ns pulses
with a 532 nm wavelength) does not generate any significant mechanical damage
on paper (based on bending, curl and tensile properties) and the ageing stability of
paper (tested through accelerated ageing) remained unaffected by the treatment
and was comparable to that of blank un-lased paper. Furthermore, it was found
that toner ablation results in the formation of nanoparticles and emission of
mostly harmless gases that can easily be captured by employing high-efficiency
particulate arresting filters and a gas extraction system, ensuring the safety of
the process.
Regarding the cost-effectiveness of the laser solution, Leal-Ayala (2012) applied
a cost of ownership model to establish the maximum capital costs that would
allow this technology to be competitive when compared with recycled paper
(considering that a ream of Xerox Supreme Recycled paper has an average
retail price of £0.03 per sheet). Assuming that the equipment has a 7 year
lifetime, a utilization of 20 per cent, variable costs of £1408 (energy) and a
throughput rate of 50 pages per hour, it is estimated that a laser un-printer
needs to have a maximum capital cost of £16 800 to compete with the £0.03
per sheet cost of recycled paper. This is considered to be a feasible goal by
the authors as they estimate that a purpose-built laser un-printer prototype
with the performance characteristics of the QuikLaze 50ST2 used in this study
(4 ns pulses with a 532 nm wavelength) could be built for around £19 000 at the
present time.
With respect to the environmental benefit of laser removal, Leal-Ayala (2012)
estimates that unprinting could translate into CO2 -eq savings between 52 and
79 per cent when compared with paper recycling (considering that 1 tonne of

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

Unprinting toner with lasers 21

recycled paper generates between 650 and 1510 kg CO2 -eq per tonne of paper
and maximum emissions from laser un-printing could be limited to 310 kg CO2 -
eq per tonne of paper). These topics define the scope of an ongoing project in
this area and future work will further expand on these results and estimates.

7. Conclusions

The toner-print removal ability of a range of UV, visible and IR lasers working
under long and ultrashort pulses has been tested in this study. The following
conclusions were obtained:

— Toner print can be effectively removed from paper by UV, visible and IR
ultrafast laser ablation.
— Only two ultrafast lasers at 532 and 1064 nm with 10 ps pulses are capable
of leaving paper with a high-enough quality to replace new paper. The
former resulted in slight yellowing of the paper substrate while the latter
produced minor fibre bleaching and cohesion loss.
— UV radiation failed as an appropriate solution to the problem under study.
This behaviour could be attributed mostly to the creation of photochemical
degradation reactions in cellulose as a result of UV light absorption.
— Green light at 532 nm with 29 ns pulses results in ablation of the secondary
materials in paper (lignin, hemicellulose and so on). This effect is no longer
appreciated when the pulse length is reduced to 4 ns as this leads to a
higher peak power and lower energy fluence combination, which prevents
any elements in paper from being ablated. The latter represents the best
laser set-up for toner-print removal.
— IR light at 1064 nm with 40 ns pulses damaged and yellowed the paper
areas under the print but did not interact with blank paper. The yellowing
could be accredited to thermal degradation accompanied by the production
of yellow chromophores as a result of heat diffusion from the toner layer
to the paper substrate. This effect disappears when the pulse length is
reduced towards 10 ps since this leads to a much quicker ablation of the
toner layer, leaving no time for heat diffusion between toner and paper.
— Both IR and visible trials using 10 ps pulses resulted in discoloration of
the paper substrate owing to the accelerated generation of discoloured
chromophores.
— All lasers using femtosecond pulses generated great physical damage on
paper owing to the vaporization of highly localized material volumes.
— An optimum operating window for the laser ablation process has been
presented for the ablation of HP LaserJet Q1338A-AC-D black toner
on white, uncoated, wood-free 80 g m−2 Canon copy paper. Future work
should focus on expanding this window by testing a wider range of
toner–paper combinations.
— Two ablation models taken from the literature have been used to
predict the ablation depths obtained when removing HP LaserJet
Q1338A-AC-D black toner from paper. The models compared favourably
with experimental measurements, suggesting that the toner-removal
process can be controlled.

Proc. R. Soc. A
Downloaded from http://rspa.royalsocietypublishing.org/ on September 26, 2018

22 D. R. Leal-Ayala et al.

We would like to thank Johannes Heberle, Peter Bechtold and Ulf Quentin at the Lehrstuhl für
Photonische Technologien for their assistance in setting up some of the experiments reported in
this article.

References
Carter, H. A. 1989 Chemistry in the comics. J. Chem. Educ. 66, 883–886. (doi:10.1021/ed066p883)
CEPI. 2006 Europe global champion in paper recycling: paper industries meet ambitious target.
Brussels, Belgium: Confederation of European Paper Industries.
CEPI. 2009 CEPI sustainability report 2009. Brussels, Belgium: Confederation of European Paper
Industries.
Counsell, T. A. M. & Allwood, J. M. 2006 Desktop paper recycling: a survey of novel technologies
that might recycle office paper within the office. J. Mater. Process. Technol. 173, 111–123.
(doi:10.1016/j.jmatprotec.2005.11.017)
Counsell, T. A. M. & Allwood, J. M. 2007 Reducing climate change gas emissions by cutting out
stages in the life cycle of office paper. Resour. Conserv. Recycl. 49, 340–352. (doi:10.1016/j.res
conrec.2006.03.018)
Daniels, V. D. 1996 The chemistry of paper conservation. Chem. Soc. Rev. 25, 179–186.
(doi:10.1039/cs9962500179)
Hekkert, M. P., van den Reek, J., Worrell, E. & Turkenburg, W. C. 2002 The impact of material
efficient end-use technologies on paper use and carbon emissions. Resour. Conserv. Recycl. 36,
241–266. (doi:10.1016/S0921-3449(02)00081-2)
HP. 2004 Q1338A materials safety data sheet: HP color LaserJet black print cartridge. Bracknell,
UK: Hewlett-Packard Ltd.
Ihori, H., Inagawa, Y., Ito, N., Fujii, M. & Ninomiya, H. 2009 Study of the conditions of irradiating
laser for removal of toner from used paper. IEEJ T. Fund. Mat. 129, 205–210. (doi:10.1541/iee
jfms.129.205)
Kaminska, A., Sawczak, M., Cieplinski, M., Sliwinski, G. & Kosmowski, B. 2004 Colorimetric study
of the post-processing effect due to pulsed laser cleaning of paper. Opt. Appl. 34, 121–132.
Kolar, J., Strlic, M., Müller-Hess, D., Gruber, A., Troschke, K., Pentzien, S. & Kautek, W. 2000
Near-UV and visible pulsed laser interaction with paper. J. Cult. Heritage. 1, S221–S224.
(doi:10.1016/S1296-2074(00)00149-7)
Kolar, J., Strlic, M. & Marincek, M. 2002 IR pulsed laser light interaction with soiled cellulose and
paper. Appl. Phys. A-Mater. 75, 673–676. (doi:10.1007/s003390201309)
Leal-Ayala, D. R. 2012 Paper re-use: toner-print removal by laser ablation. PhD Thesis, University
of Cambridge, UK.
Leal-Ayala, D. R., Allwood, J. M. & Counsell, T. A. M. 2011 Paper un-printing: using
lasers to remove toner-print in order to reuse office paper. Appl. Phys. A. 105, 801–818.
(doi:10.1007/s00339-011-6654-z)
Liu, X., Du, D. & Morou, G. 1997 Laser ablation and micromachining with ultrashort laser pulses.
IEEE J. Quantum. Elect. 33, 1706–1716. (doi:10.1109/3.631270)
Preuss, S., Demchuk, A. & Stuke, M. 1995 Subpicosecond UV laser-ablation of metals. Appl. Phys.
A. 61, 33–37. (doi:10.1007/BF01538207)
Ready, J. F. 1965 Effects due to absorption of laser radiation. J. Appl. Phys. 36, 462–468.
(doi:10.1063/1.1714012)
Schmidt, M. J. J., Low, D. K. Y. & Li, L. 2002 Laser ablation of a B4 C-polysiloxane composite.
Appl. Surf. Sci. 186, 271–275. (doi:10.1016/S0169-4332(01)00629-8)

Proc. R. Soc. A

Anda mungkin juga menyukai