Overview of Forming LITERATURE, 1990-2000: Bo Norman Daniel Söderberg
Overview of Forming LITERATURE, 1990-2000: Bo Norman Daniel Söderberg
In
The science of papermaking, Trans. of the XIIth Fund. Res. Symp. Oxford, 2001, (C.F. Baker, ed.),
pp 431–558, FRC, Manchester, 2018. DOI: 10.15376/frc.2001.1.431.
OVERVIEW OF FORMING
LITERATURE, 1990–2000
ABSTRACT
TABLE OF CONTENTS
1 Background 435
2 Paper structure characterisation 436
2.1 Power spectra 437
2.2 Specific perimeter 439
2.3 Wavelets 439
2.4 Formation 441
2.5 Medium-scale variations 446
2.6 Fibre orientation 447
3 Fibre suspensions 450
3.1 Fibre suspension modelling 450
3.2 Fibre flocculation evaluation 453
3.3 Velocity and turbulence measurements 459
4 Headboxes 462
4.1 Tapered header and CD control 462
4.2 Tube bank 465
4.3 Headbox nozzle 467
4.4 Headbox jet and streaks 480
5 Wire section designs 487
5.1 Roll-blade dewatering 487
5.2 Multi-ply board forming 496
6 Web forming 496
6.1 Random sheets 496
6.2 Self-healing effects 498
6.3 Formation improving mechanisms 501
6.4 Fibre deposition on wire 502
6.5 Wet web resistance 505
7 Dewatering processes 508
7.1 Jet impingement 508
7.2 Fourdrinier dewatering 508
7.3 Blade dewatering 510
7.4 Roll dewatering 517
7.5 Vacuum dewatering 521
8 Laboratory forming 523
8.1 Sheet forming 523
8.2 Pilot machines 528
1 BACKGROUND
meters – characterized by the fact that they never actually intend to directly
measure concentration.
The degree of fibre orientation anisotropy and fibre orientation misalignment
is increasingly important. Regarding terminology, fibre orientation to many
seems to specifically denote the misalignment angle. It is recommended that
anisotropy or misalignment respectively be added to avoid misunderstandings.
What used to be a wire now increasingly is denoted a forming fabric. This is
to emphasize that what was originally a two-shaft bronze material weave
is now a much more sophisticated design based on polymer threads. Cor-
respondingly, what used to be a felt is nowadays a press fabric. In this
overview the old, simpler terms will still be used to some extent. Thus, e.g.,
twin-wire, wire section, wire mark and felt mark will be the preferred terms. In
this context it does not seem quite compatible that a fabric generates a wire
mark. On the other hand, it could not generate a fabric mark, since a press
fabric might equally well have generated that.
The term roll-blade former will often be used for all designs where a roll is
followed by some kind of blade arrangement. In principle, this term could
thus also include the roll-adjustable blade design, in which the forming roll is
followed by a combination of fixed and movable blades on the two sides of
the wires.
The use of different material mixtures on different levels in a product as
now is increasingly applied to different products can be achieved using two
basically different principles. The tradition is to manufacture board in several
plies, while e.g., tissue products are increasingly formed in different layers.
The multi-layer products are formed using a single headbox, with different
component mixtures in different headbox layers, while a multi-ply product
consists of plies formed from separate headboxes.
For readers, not familiar with the basics of forming, it could be advisable
to initially study a textbook covering the area, such as Papermaking Part 1,
Stock Preparation and Wet End in the new book series Papermaking Science
and Technology [3].
• Small scale covers the range of 0.1 to 40 mm, in which range fibre floccula-
tion and hydrodynamic conditions during forming are the main causes for
irregularity.
• Medium scale covers the range of 40 mm to 10 m and is affected mainly by
instabilities in headbox flow (partly generated in the approach flow system)
and wire section dewatering.
• Macro scale variations are those in excess of 10 m and are caused mainly by
the variability in the incoming thick stock. The variations are then mainly in
the machine direction. Such variations will not be treated in this overview.
Figure 1 1-D sample variations (top), hypothetical waveform and wave displacement
for evaluation of Fourier spectrum (middle) and Power spectrum (bottom).
parts by a vertical line at the mean wavelength. Mean floc size will then
amount to half that mean wavelength; see e.g., [4].
For practical reasons, the wavelength power spectrum is often presented
using logarithmic scales on the spectral density axis (and sometimes also
on the wavelength axis). When two spectra are compared, a certain vertical
difference between the spectra then corresponds to a specified ratio between
the spectral densities, independently of the absolute level.
2.3 Wavelets
Wavelet methods give a more complete description of local variations in
comparison with power spectra. While the power spectrum gives an average
of variations of different scales over the whole sample area, wavelets also
specify the location of specific variations. In the power spectrum, variations
are described based on comparisons with sine wave shaped curves, while
with the wavelet technique the choice of curve shape is free (even if some
mathematical restrictions exist). Two examples are given in Figure 2.
An application of wavelet technology (Morlet) is demonstrated for the
Figure 2 Left: Mexican-hat wavelet, Right: Morlet wavelet (sinusoidal signal weighted
with a Gaussian distribution), real part (solid line) and imaginary part (dashed line).
signal sin(x2) as function of distance x, see Figure 3a. The power spectrum of
this curve is shown in Figure 3b, and demonstrates high spectral density at
short wavelengths, and a gradual decrease towards longer wavelengths.
In the image map from the wavelet analysis, Figure 3c, the wavelength is
plotted as a function of distance, and it is apparent that the wavelength
decreases towards longer distances. This information is not available in a
power spectrum.
In Figure 3d the same result is presented using contour lines.
Figure 3 a) sin(x2) signal, b) Power spectral density, c) Image map of wavelet (Morlet)
transform and d) Contour plot of wavelet transform with exact wavelength variation
(dashed line) [7].
Use of wavelet techniques has been discussed by Keller et al. [5,6] and
applied by Söderberg [7] (see further below).
2.4 Formation
The formation of a paper sheet, i.e., the local grammage variations up to
about 40 mm wavelength, is determined by the fibre distribution in the plane
of the sheet and has a great influence on many sheet properties; it is therefore
desirable to be able to quantify this distribution. The geometrical resolution
of the basis weight measurement is of decisive importance for the amount of
variations recorded. The smaller the measurement area is, the more small-
scale variations can be detected and the larger the total variations recorded.
In industry, formation evaluation is often done using light transmission
methods. Besides local grammage variations, this method is however also
sensitive to other parameters, primarily variations in local light scattering
coefficient. Beta radiation is a good alternative to be used in evaluating
grammage variations.
Formation characterisation
The Formation number F is often used to denote the coefficient of variation of
local basis weight, that is, standard deviation divided by mean basis weight
and it is often expressed as a percentage [1]. One exception was the use by
Dodson of the same term to denote the variance of local grammage of a
paper sample, normalised with that of a sheet with random fibre distribution
[18]. Although physically sound, this use of the term is unpractical, since
knowledge of the corresponding random fibre distribution is usually missing.
It would require complete knowledge of dimension and coarseness values for
all the individual fibres.
Normalisation with the formation of a standard laboratory sheet made
from the same furnish has also been used be several researchers, e.g., Lloyd
[19].
Regarding normalising the formation values for differing mean grammage,
a method based on statistical means has been used [1]:
tions sometimes can reduce the number of large flocs by breaking them apart,
and thus move them to a smaller floc size range. This may not always show up
in a single formation number.
In the STFI method, formation is described by its wavelength spectrum.
The information in a complete formation spectrum can be simplified by
integration within different scale ranges. In the STFI method, a small-scale
wavelength range of 0.3 to 3 mm and a large-scale range of 3 to 30 mm are
used, and together they make up the total formation number F.
In the Ambertec meter, the limited sample size sometimes requires
measurements on several samples to give accurate values of the standard
deviation. The normalized Ambertec formation value with dimension
“(g/m2)0.5” (not easy to understand) can actually be interpreted as the dimen-
sionless normalized formation number F according to Equation 1, using a
normalization basis weight of 1 g/m2. It is thus possible to compare an
Ambertec to a STFI value by multiplication with 600.5 ≈ 7.7. It should how-
ever be remembered that the Ambertec meter suppresses all variations
smaller than 1 mm, so its formation value mainly correlates with the STFI
large-scale value.
Dodson et al. used the variance against zone size to characterise the scale
distribution in basis weight variations [20]. This is defined as the integral of
the wavelength spectrum from the specified zone size to some upper limit.
This makes it hard to evaluate differences in small-scale variations between
two samples, since the influence of the dominating variance at larger zone
sizes will always dominate.
An alternative method of evaluating formation scale is by the specific per-
imeter as promoted by Jordan et al. [21].
Using wavelets to characterise formation, it is possible to describe the
distribution of floc sizes within different parts of a paper sample, not only
give the average over the whole area. This method has been applied by Keller
et al. [5].
Periodic marking
While formation measures mainly quantify the degree of fibre flocculation in
a paper sample, different kinds of periodic marking is also of interest. Such
marking is mainly generated by wire patterns, but sometimes also by the felts.
In a one-dimensional formation spectrum, wire marks will show up as peaks
in the spectrum. It will then be possible to rank samples depending on the
magnitude of such peaks.
A more complete way of describing periodic marks is to first calculate
the two-dimensional frequency power spectrum, in which periodic variations
These points can then be inversely transformed back into the original x-y
plane, where the geometric pattern of the generating wire and/or felt will
show up, see e.g., [22,23].
Optical measurements
The term “formation index” is reserved for optical formation meters. Such
meters have the drawback that they do not give absolute values of basis
weight variations, due to their sensitivity to variations in light scattering
coefficient.
This has been demonstrated repeatedly, recently by Duffy et al. [24]. Two
light transmission meters were compared with the Ambertec and STFI
methods (beta ray transmission). The light transmission methods gave reverse
results from the beta ray methods when e.g., grammage was changed in
laboratory forming.
Luner et al. [16] concluded that for newsprint samples, electrography gave
higher spatial resolution, shorter exposure time and wider basis weight range
than β-radiography. Light transmission gave the poorest spatial resolution
and correlation with mass. Soft X-radiography gave the highest spatial
resolution but the poorest contrast.
Bernié and Douglas [25] compared light transmission and beta ray trans-
mission and came to the conclusion that light transmission can be used
to measure formation. This was however based on absolute calibrations and
was applied to laboratory sheets only. For practical purposes, this is of less
interest.
Bouydain et al. used uncalibrated light transmission images in combin-
ation with wavelet analysis to study paper formation [26].
However, optical methods due to their simplicity can still be valuable, for
instance to follow changes on a specific machine at constant furnish, basis
weight and calendaring conditions. Absolute comparisons between paper
samples from different machines should be avoided.
Two of the more commonly used optical formation testers in North
America are the NUI (Non Uniformity Index) meter [27] and the M/K-meter
[28]. The NUI-meter analyses light transmitted through a rotating paper
sample, and a total variation number is given as a formation index. With the
M/K-meter, on the other hand, a high value corresponds to good formation.
Recently also the Kajaani optical formation tester has been introduced and,
like in the M/K-meter, the formation index is higher for a more even paper
sample.
If the aim of an optical formation meter is to resemble perceived look-
through formation the wavelength distribution of the light analysed should
be matched with the colour response of the human eye [29].
Lloyd [19,30] introduced the parameter “surface ply variation”, SPV, to
characterise the final mixing in paper samples manufactured from a three-
ply headbox. A light blue colour was added to the middle ply, to make it
possible to optically evaluate the degree of mixing between centre and outer
layers. 87 × 174 mm2 samples were analysed using a colour scanner at a
resolution of 300 dpi. The cyan part of the image was separated. This
image was inverted, i.e., a negative taken, because the cyan represented an
absence of the white surface layer, and it was the absence of the surface
layer that was of most interest. The resulting image was a grey-scale TIFF
image. The image was analysed with a procedure similar to that used for the
evaluation of the small-scale and large-scale STFI formation values; see
further below.
On-line measurements
Today optical CCD-techniques can be applied to both laboratory and on-line
evaluation. One example is the formation meter developed at CTP [31]. Using
a CCD camera a sample size of 120 × 120 mm2 is scanned with 0.25 mm
resolution. From this a total formation index can be evaluated as well as flocs
in the size classes 1, 2, 3, 6, 10 and 16 mm. It is also possible to evaluate shape
(MD/CD) as well as angle of orientation of flocs, see Figure 5. Means are
also provided to filter out and present periodic wire markings. In on-line
Figure 5 Presentation of results from the CTP on-line formation analyser [31].
Figure 6 Grey-scale picture (original in colour) of basis weight of full width paper
web. Newsprint quality, with each level curve indicating a deviation of 1 g/m2 [32]
(Courtesy STORA, original in colour).
is even possible to on-line follow the patterns of the narrow wire shower
marks using this technique.
Honeywell-Measurex developed a method for detecting water weight in the
wire section, aiming at following the CD and MD dewatering profiles [35].
Non-scanning sensor technology was used, based on the measurement of
effective electrical properties of water between sensor elements mounted
below the wire. It is yet uncertain if it will develop into commercial equipment.
Laboratory evaluation
Hasuike et al. [39] analysed fibre orientation anisotropy after splitting a
sample into eight layers using adhesive tape; see Figure 7. Fibre orientation
anisotropy in the layered samples was evaluated from manual counting of the
number of fibre crossings with a 5 mm-long MD and CD test line respectively,
within sample areas of 30 × 30 mm2.
They further analysed the three-dimensional structure of paper samples
[40,41]. Fibre entanglement in the thickness direction was evaluated by
embedding a sample in epoxy resin, cutting it into 4μm thick sections (2 mm
wide) and following the positions in the thickness direction of individual
fibres along its length, through 150 sections.
Hasuike et al. [42] also studied 3-D floc structure in 45 × 95 mm2 paper
samples using tape-layering techniques. Floc distribution in each layer was
evaluated optically, and the correlation between floc positions in adjacent
layers was utilised to quantify floc dimensions in the thickness direction. A
limitation of this method is that tape delamination does not strictly split the
paper samples into even layers, since some flocs tend to resist splitting.
Erkkilä et al. also used a method based on tape splitting [43]. A 60 g/m2
sample was split into ca 6–8 layers, on which 30 × 30 mm2 samples were
analysed in transmitted light using a CCD-camera with a resolution of
640 × 640 pixels. An image analysis program was developed to evaluate the
orientation distribution of fibre segments in the images. The results were
presented in polar diagrams where the orientation distribution has an elliptic-
like shape. Anisotropy was evaluated from the eccentricity b/a, and mis-
alignment angle as the deviation of the main axis from MD; see Figure 8.
Local fibre layer orientation was also studied, on sub areas of 3 × 3 mm2,
and presented in vector form.
Jansson [44] used another tape quality, and separated an 80 g/m2 sample
into 20 layers, which he imaged using a desktop scanner with a resolution of
600 × 600 dpi. He analysed the images of the size 40 × 40 mm2 using the
program developed by Erkkilä.
Lloyd and Chalmers used similar techniques to study the effects of sheet
structure on paper curl and cockle [45].
Scharcanski and Dodson [46] demonstrated an image gradient evaluation
technique to quantify anisotropy from a beta-radiograph or a light trans-
mission image.
Thorpe [47] used a “new method” to split a copy paper into 13 layers. The
black coloured layers were mounted in 35 mm slide frames and imaged with
a CCD camera attached to a microscope. Fibre orientating data were evalu-
ated using the Hough transform. He analysed samples of size 20 × 25 mm2,
and each picture was divided into 5 × 5 mm2 segments for local orientation
analysis. He could evaluate the occurrence of orientation streaks on the wire
side of fourdrinier formed paper.
Parker et al. [48] studied local fibre orientation at different levels in the
thickness direction using confocal laser scanning microscopy and application
of the Hough transform. The thickness resolution was ≈ 3μm and the depth
was limited to 80μm. The scanning area for each sample was limited to
650 × 650 μm2.
3 FIBRE SUSPENSIONS
2
2 L
N = cv
3 d 冢冣 (3)
Table 1 is a summary of the findings using the turbulence decay cell described
above. In standard hand sheet forming, the crowding factor approximately
equals unity.
Wikström et al. [55,56] with the main aim of improving pulp processing,
which generally means concentration ranges higher than those in paper-
making. It was e.g., suggested that fibre network yield stress was a better
measure than fibre concentration for designing pulp lines [57]. Experimental
studies of fibre-fibre friction were combined with fibre force models to predict
fibre network properties [58,59].
Experiments by Bennington [60] , based on modelling of chemical reactiv-
ity indicated that in a 3% fibre suspension, 95% of the energy dissipation
takes place by direct friction between the fibres and thus only 5% in the fluid
itself.
Huhtanen [62] studied the non-Newtonian properties of fibre suspensions,
without taking fibre flocculation specifically into account. He made rheo-
logical measurements to analyse shear viscosity of pulp suspensions at mod-
erate shear rates and small deformations. He also applied a commercial code
to describe the non-Newtonian effects on pipe flow of pulp suspensions.
Klingenberg et al. have an extensive program to study flowing suspensions
of rigid and flexible fibres, modelled as chains of prolate spheroids connected
through elastic ball and socket joints [63]. Attractive forces between fibres can
form flocs, but the properties of such flocs are different from flocs formed
from elastic interlocking and inter-fiber friction. Simulations suggest that
fibre shape is important and that effective formation aids in paper-
making should reduce inter-fibre friction under normal force loadings of
1–10mN [61,64,65]. Examples of the influence of friction coefficient on the
flocculation in shear flow shown in Figure 9.
Shah et al. [66] studied the flow patterns generated by a half Δ-wing in a
water tunnel, specifically the fluid stress fields. Applying these fields on fibre
suspension models, they came to the conclusion that the local stresses were
large enough to break apart fibre flocs.
Figure 10 Left: Concept for grid pulled through a vertical, narrow channel. Right:
State of flocculation for Douglas fir, 0.5% , Crowding Factor N = 75 [52].
down through the fibre suspension in the cell. After turbulence decay (app.
15s) a flocculation image was photographically recorded in transmitted light.
From this image, contrast intensity and floc scale was evaluated.
Kallmes et al. [68] demonstrated equipment for on-line analysis of mix
flocculation. Mix was extracted from the headbox into a 13 mm-wide chan-
nel. The light transmission variations of a narrow light beam directed
through the fibre suspension gave a flocculation value. A comparison was
made with the inherent flocculation tendency of the mix, as recorded after the
passage of a dispersing 90° flow elbow, as well as with a formation index
recorded with an on-line, light beam transmission meter. Preliminary test on
different paper machines showed promising correlations between flocculation
and formation values in response to e.g., changes in chemicals addition.
Zhao and Kerekes [69] used the above-mentioned cell with decaying turbu-
lence to study fibre flocculation at different suspending liquid viscosities. One
way of presenting the results is to find the “critical crowding factor”, below
which no flocculation can be detected. In pure water (viscosity 1 mPas) the
critical crowding factor for long fibre pulp is around 10, but increases linearly
to 70 when sugar is added to increase liquid viscosity to 20 mPas. This is
explained by reduced turbulent motion, decreased initial turbulence and
faster turbulence decay, and higher shear forces on the fibres.
Duffy et al. [70] made a study of pipe flow of fibres suspended in liquids
of different viscosities. The numerous flow regimes normally observed for
friction loss curves in pipe flow were not present for viscous suspensions.
Drag reduction is also observed in viscous suspensions but it onsets at higher
velocities. Increasing viscosity leads to reduced floc size in pipe flow. Overall,
it is suggested that increasing viscosity has the same effect as decreasing fibre
concentration.
Giro et al. studied the effects of adding CMC dispersants to high concen-
tration pulp slurries ( ≈ 30%) [71]. The torque requirement in a bowl mixer
was used as a measure of pulp dispersibility. It was concluded that increasing
CMC molecular weight had a positive effect, and pulp dispersion was more
closely related to the amount of CMC adsorbed than to the amount in solu-
tion. This indicates that at this high concentration level, fibre friction seems to
be the determining factor, and not fluid viscosity.
Kerekes and Schell [72] , also used the decaying turbulence cell to study the
effects of fibre length and coarseness on flocculation. Their main findings are:
Figure 11 Drawing of the turbulence generator and camera system for the floc-
measuring device (not to scale). Turbulence generator channel widths 15 mm and
length 80 mm. Lamellae length 270 mm and thickness 1 mm [73]
means that there is still uncertainty whether or not the higher Reynolds num-
ber at realistic dimensions will generate better mixing.
Ringnér and Rasmuson [84] applied X-ray tomography to study 3D fibre
distribution under stationary conditions. Their main interest was to investi-
gate inter and intra floc concentrations in a range above 3%, to improve the
understanding of pulp treatment processes.
Kellomäki et al. [85] used a fast CCD-camera to detect light transmission
through a 30 × 40 mm2 cross sectional rectangular channel. They compen-
sated for unevenness in the background illumination and relied on the
Lambert-Beer law to calculate fibre concentration as a function on the
amount of transmitted light. They evaluated floc size at a grey scale level of
average intensity, so that 50% of the image was always interpreted as fibre
flocs. They assumed cylindrically shaped flocs, and calculated the effective
dimensionless floc volume Vf* from the equation
2
LmxLmy
V f* = , (4)
l f3
where Lmx is the mean floc size in the flow direction, Lmy is the mean floc size
in the cross direction and lf is the mean fibre length.
Karema et al. [86] studied the transient fluidisation of fibre suspensions
after a step diffuser, using an optical flocculation analysis method described
above. They evaluated the dimensionless floc volume Vf* (see Equation 4) as a
function of distance/time after the step at different levels of expansion head
loss, which was obtained by different flow velocities; see Figure 15.
Short fibre and long fibre pulps were used. The results indicate that the
growth of the dimensionless floc volume Vf* during re-flocculation follows a
common power law with respect to time for the pulp types studied. At later
times, floc size saturates at a constant value.
Ultrasound
Karema et al. used ultrasound to measure fibre turbulence in the thickness
direction of a 30 mm thick, plane channel [86]. The instantaneous fibre phase
Figure 15. Dimensionless floc volume Vf* as a function of time tc elapsed from step
expansion at different head loss hf during expansion [86].
NMR tomography
Fibre suspension flow was studied by Li et al. using NMR imaging tech-
niques. Pipe flow profiles at flow velocities up to 1 m/s were studied [87].
Various flow patterns from plug flow to complex flow were observed for
hardwood suspensions of ca 0.5% fibre concentration. Only plug flow was
observed at 0.9% for a softwood pulp suspension. Studies were also made of
pulp flow through an abrupt contraction [88], both with water flow and pulp
suspension flow. For the pulp suspension, they observed plug flow in the
regions both far upstream and far downstream of the contraction. Just
4 HEADBOXES
distributor. The water is led into the fittings where the flow leaves the tank. By
this local CD addition of water, the local mix concentration can be con-
trolled. Since the basis weight is changed already inside the headbox it is no
longer necessary to use the slice lip to control the grammage profile. Hence
fibre orientation misalignment due to the induced cross-flow can be avoided
and grammage control is de-coupled from fibre orientation misalignment.
The BTF dilution control system has also been designed to be retrofitted on
existing paper machine headboxes; see Knoller et al. [95].
The concept of dilution control is today also applied in standard headbox
designs, where the stock is supplied to the headbox by a tapered header. The
first implementation was a Voith-Sulzer Module Jet headbox in Munkedal,
which was put into operation in January 1994 [96,97,98]. The dilution control
concept has been widely adapted and there are several technical solutions
used to add the extra water.
In Figure 18 (top) the addition of dilution water in the Voith-Sulzer Mod-
ule Jet headbox can be seen [97]. Mix and white water is fed to the headbox.
Figure 18 Three different concepts of dilution control. Top: Voith [97] , Bottom, left:
Beloit [99]. Bottom, right: Valmet [91].
Motorised valves control the addition of white water, which is blended with
the mix in a mixing chamber located in a first, single-row tube bank. An
orifice is located after the mixing chamber and the unit (valve, mixing cham-
ber and orifice) is designed so that the volume flow through the orifice
(mix + dilution water) is kept constant when the amount of dilution is
changed. A good blending of the additional white water and mix is obtained
by injecting the dilution water at a specific angle into the mixing chamber.
Figure 18 (bottom, left) shows the technique used by Beloit in the Concept
IV-MH headbox [99,100]. The white water is added already in the tapered
header. Since it is added before the entrance into the tubes the flow rate into
the tubes is only given by the pressure distribution in the tapered header.
Hence it remains practically unaffected by the addition of extra water. The
direction of the added water into the manifold assures a good mixing.
In Valmet’s Sym-Flo D headbox, there are two tube banks separated by a
stilling chamber and the additional water is injected directly into the tubes in
the first tube bank, at a step change in pipe diameter; see Figure 18 (bottom,
right). The constant pressure in the stilling chamber then assures a constant
flow rate across the headbox nozzle; see Nyberg and Malashenko [101].
Dilution control is a powerful technology to obtain good CD control of
grammage and fibre orientation misalignment, and represents a leap forward
in paper process technology. It is widely used on paper machines throughout
the world. Many investigations have been made to study the application of
dilution control systems in different applications [102,103,104,105,106,107].
Lee and Pantaleo [108] performed CFD simulations aimed at investigating
the influence of headbox flow features on fibre orientation.
In order to achieve the most uniform grammage and fibre orientation CD-
profiles, Hämäläinen and Tarvainen [109] introduced optimisation methods
coupling numerical simulations (CFD) of headbox flow and paper properties
such as grammage and fibre orientation misalignment. They illustrated the
method by two industrially applicable examples. These were the shape of the
tapered header and grammage control both by slice bar position and dilution.
In order to perform an optimisation, a cost function had to be defined. For
the case of the tapered header the cost function should capture the departure
from a uniform flow velocity CD profile. This can be i.e., the standard devi-
ation of the flow velocity in CD. For the case of slice and dilution control the
cost function could be based on grammage and/or alignment non-uniformity.
The solution to the problem was obtained by standard optimisation
techniques.
water and also studied the flow conditions in the tubes with particle
visualisation.
The results can be seen in Figure 19 where the pressure drop can be seen as
a function of velocity through the tubes. The different geometries are
included in the figure. As can be seen the highest obtained pressure drop is
four times as high as the lowest.
Hauptmann et al. [111] discussed the effect of turbulence generation by
backward facing steps and wakes, and made experiments using hot-film
anemometry in water. They looked at the generation and decay of turbulence
and the turbulent length-scales. Comparisons of the results were made with
visualisations of the flow behaviour of the jet and the conditions on the
forming table.
Shariati et al. [112] numerically solved the flow through a manifold, tube
bank, contraction and slice. The tubes in the tube bank had a somewhat older
design with a tapered diffuser instead of a step diffuser. The results clearly
showed the effect of re-circulation, which was generated by flow separation in
the diffuser. Since the complete flow was solved, the effect of non-uniform
flow conditions showed a strong influence in the behaviour of the flow inside
the tubes depending on their position in the headbox.
Due to the rapid re-direction of the flow when it enters the tube from
the manifold, flow separation is also likely to occur at the inlet end on the
upstream side of the tube, immediately after the tube entrance. Since the
Figure 19 Tube pressure drop as function of inlet velocity for six different tube
designs [110].
step-wise increase in area with its associated large-scale separation and re-
circulation will generate turbulence, it will have an influence on the dispersion
on fibre flocs. Also, the strong acceleration of the entrance flow into the tubes
can possibly contribute to the dispersion of flocs as well as to aligning fibres
in the flow direction.
However, the strong turbulence generation in the step diffuser will eventu-
ally also re-generate flocs through the strong mixing.
Karema et al. [86] presented results regarding the effect of transient fluid-
isation in channel flow due to a step increase in area, which is closely related
to the effect of step diffusers in the tube package. Their experiments are
described in section Fibre suspensions. The results show a flocculation min-
imum some distance downstream of the steps, after which the flocculation
increases; see Figure 15.
In a typical design, the manifold flow towards an individual hole is not
perfectly symmetrical, which can generate a strong swirl into the tube. With
the flow rates used in paper manufacturing this could give a considerable
effect and create strong vortices in a tube bank. Even if this can occur gener-
ically, the effect has been applied wilfully, by generating strong swirls inside
the tubes in order to promote isotropy in the final paper sheet, Aidun
[113]. This has been later patented [114] and is discussed further below; see
Figure 23.
speed was changed. Zero mix-to-wire speed difference was defined at the wire
speed giving minimum fibre orientation anisotropy. Due to the minimum
amount of shear introduced during these dewatering conditions, fibre orien-
tation anisotropy in the paper produced should be a good approximation to
that already existing in the jet.
At low and medium headbox nozzle contraction ratios (traditional for
hydraulic headboxes), fibre orientation anisotropy in the paper produced
behaved in the traditional way, with some increase of anisotropy at the min-
imum point, at increased contraction ratio; see Figure 20. At high nozzle
contraction, however, very high fibre orientation anisotropy values were
obtained even at zero speed difference. This indicates that fibre orientation
anisotropy was high already in the jet. The logical reason for this is the
strongly elongational nozzle flow. It was also found that paper formation was
improved with higher nozzle contraction ratio, which indicates that fibre
flocs were broken apart in the elongational flow – see further in a later
section.
As stated above, headbox jet quality improves with increasing nozzle con-
traction ratio. It would then be a natural choice to use a high nozzle contrac-
tion ratio in all headboxes. However, since most paper and board products
require a low fibre orientation anisotropy, a high contraction ratio can only
be utilised for products allowing a high MD/CD ratio, such as printing papers
containing mechanical pulp.
Ullmar [118,119] studied the fibre orientation anisotropy effects in a trans-
parent headbox nozzle. A flat channel extended the slice, to facilitate optical
measurements on the emerging jet; see Figure 21. A stroboscope and a video
camera were used to record the fibres in the jet, and image analysis was
applied to analyse fibre orientation characteristics.
Nozzle contraction ratio was varied within the range 17–50. Due to the
opaqueness of fibre suspensions, it was not possible to view individual fibres
more than a few mm inside of the jet surface. Therefore the majority of the
experiments were performed at very low fibre concentrations. Coloured
polymer fibres (comparatively straight) as well as cellulose fibres (more curly)
were used. Some of the conclusions are listed below:
Figure 21 Transparent headbox for fibre orientation studies in the headbox jet; to
the upper right: video camera and stroboscope [118].
Table 2 shows the effects when two different contraction ratios, CR, and
two different nozzle lengths, L, are applied. It is assumed that the flow vel-
ocity at the slice is constant, which also means that the volumetric flow rate is
constant.
Given a large contraction ratio and a short nozzle the acceleration is high-
est, which would give the strongest strain on the suspension. The opposite
would be a small contraction ratio with a long nozzle, which would give a low
acceleration and weak strain. However, the longest time is spent in the elon-
gational flow field for the long nozzle with a large contraction ratio and the
shortest for the short nozzle with a small contraction ratio. Hence, if the time
spent in the contraction is of importance for the final paper sheet it is really a
two-parameter problem.
Asplund [120] used the headbox in Figure 21 to study the fibre orientation
anisotropy at different levels in the jet. Instead of a stroboscope, a ca 1 mm
thick laser light sheet was used as a light source. In Figure 22, results are
shown with a centrally inserted vane in the nozzle, ending 30 mm upstream of
the nozzle end. Fibre orientation anisotropy is significantly lower in the upper
lip boundary layer than outside of it. The thin vane produces approximately
the same decrease in anisotropy at jet centre as the boundary layer effect,
while the thicker vanes generate further disperging fluid motions.
Aidun [114] suggested a method to decrease fibre orientation anisotropy
after the tube package feeding the headbox nozzle; see Figure 23. The left
illustration shows an insert, which is mounted inside each tube in the tube
bank. This insert will generate a strong and controlled swirl. The swirls that
are generated by the different tubes are not all rotating in the same direction.
The amount of tubes generating clockwise vortices is the same as the number
that generates counter-clockwise vortices and the two types of tubes are
arranged in specially designed patterns. An example can be seen in the right
illustration where a staggered pattern is used. To obtain an overall design so
that all swirls cancel each other before leaving the headbox nozzle will be a
delicate task.
Figure 22 Fibre orientation anisotropy in the upper half of a 15 mm-thick jet. Three
alternative vanes inserted: thin, thick with pointed tip and thick with blunt tip [120].
Figure 23 Swirl generation in feed pipes (Left) and rotation directions in adjacent
feed pipes (Right) according to patent by Aidun [114].
Figure 25 Mix surface roughness profiles with and without headbox nozzle vane [122].
the inlet conditions for the 3-D case are much more physically relevant.
The result of the 3-D simulations was in good agreement with experiments
regarding the behaviour in the contracting nozzle. However, by comparison
of different figures in the article regarding the behaviour of turbulent kinetic
energy, it is clear that a comparison with experimental results was not per-
formed for the end of the nozzle contraction, where large differences are to be
expected. This is discussed below.
Farrington also investigated the effect of vanes in a nozzle; see Figure 26.
The graph shows the behaviour of turbulent kinetic energy and turbulent
macro scale with and without vane. The kinetic energy behaves similarly in
the two cases throughout the nozzle until the position where the vane ends,
where a strong increase is noticed. The vane tip generates turbulence, which
however decays before nozzle outlet. The vane clearly has a restricting effect
on the turbulence macro scale.
Hua et al. [124] numerically solved the complete flow in a generic headbox
using the standard k-ε model. The result showed clear structures after the
tube bank. These gradually disappeared towards the slice. Also investigations
Figure 26 CFD calculations of flow characteristics with and without vanes (DIV/
NO DIV) in headbox nozzle; effects on turbulence scale (SCALE) and turbulent
kinetic energy (TKE) along the nozzle, X(M) [123].
velocity, i.e., Urms, normalised with the value at the beginning of the contrac-
tion. Table 3 explains the different plots.
Designation Method
Figure 28 a) 3-layer headbox; effects of relative vane length and mix-to-wire speed
on SPV (MD), b) small scale and c) large scale [129].
variations and large-scale variations were evaluated. Since these are spa-
tial variations, the direction in which they are measured is of importance
(MD or CD).
The effects of relative vane length and mix-to-wire speed difference on SPV
are shown in Figures 28b (small scale) and c (large scale). Relative vane length
is the distance from the vane tip to the slice, with a negative value when the
vane ends inside the nozzle. Longer vanes seem to give lower SPV values both
for the small and the large scales. However, tape splitting of the sheets showed
that for the two longest vanes (+ 20mm and + 100mm) the flocs in all of the
layers consisted of a mixture of white and blue fibres. For the vanes that
ended inside the nozzle (primarily −95mm and −250mm) there was instead a
clear differentiation between blue and white flocs. Hence, by changing the
vane length, an optimum can be identified where the tip of the vane ends right
before the slice lip, where SPV has the lowest value, without “total” mixing.
Three headbox slice geometries in combination with different vane lengths
were also tested [30]. A parrot’s beak slice lip was compared to a linearly
converging slice lip and a plane channel slice lip. With the vanes ending inside
the nozzle, the parrot’s beak appeared to reduce layer mixing, while the
parallel channel appeared to increase it. However, the type of slice lip had
little effect on layer mixing if the vanes ended outside of the headbox; vane
tip vortices then seemed to be the dominating source of mixing. The parrot’s
beak also improved the small-scale formation.
Finally, Lloyd and Norman also studied the effect of vane shape [130]. In
one set of experiments, a 5 mm-thick vane was followed by a 0.5 mm thin-vane
of different lengths, 10–40 mm. The step height was thus 2.25 mm. In another
set of experiments, the steps were varied from 0.5 to 1.75 mm. The aim of the
steps was to break down the boundary layers formed upstream along the
vanes, and thus reduce downstream vortex generation. However, the turbu-
lence introduced by the steps increased the layer mixing, both at floc and fibre
level. This turbulence gave lower fibre orientation anisotropy, worsened
small-scale formation and Z-toughness, but improved fracture toughness of
the formed sheets.
Lloyd summarised the work on layering in a Tekn.lic. thesis [19].
Using flow simulations Parsheh and Dahlkild [131] numerically studied the
mixing behind vanes with different tip shapes and positions. The simulations
were performed with the k-ε turbulence model. The geometry was similar to
that of Lloyd and Norman [129]. In the simulations, which were two-
dimensional, the mixing was studied by adding a passive scalar into one of
the layers. The spreading of this scalar was then studied. The shortest vane
gave the lowest mixing at a relevant distance downstream in the jet. Also, hot-
wire studies of mixing behind a vane in a contraction were performed by
Parsheh [127].
The vanes in the headbox are not straight but will deform as a function of
the flow field and the bending stiffness of the vanes. Parsheh and Dahlkild
[126] modelled this problem. They developed a quasi one-dimensional model
describing the position and shape of the vane as a function of bending stiff-
ness, vane and nozzle lengths and flow conditions at the inlet. The model was
compared with direct simulations using the commercial software CFX to
model the flow and the position and shape of the vanes. This was performed
using an iterative technique.
Söderberg [7] performed a pilot machine study of the effect of contrac-
tion on headbox jet flow, a study not specifically aimed at stratified forming.
In the study the same stratified headbox as that of Lloyd and Norman
(Figure 28) was used. The length of the nozzle and vane was kept constant
and the inlet area was varied. Each layer in the headbox was given an
individual colour. The top layer was white, the middle layer was blue and
the bottom layer was red, and the headbox was mounted at the fourdrinier
position in the EuroFEX pilot machine. The headbox jet was visualised at
several MD positions after the slice and these visualisations were compared
to the paper produced, which was scanned using an ordinary desktop
scanner.
The influence of headbox contraction ratio can be seen in Figure 30. The
top row is scanned paper sheets where the blue colour (centre layer) has been
coded as black. The bottom row contains images showing the headbox jet at
the three different contraction ratios. It is clear that the characteristics of the
jet are reflected in the structure of the paper sheet. The low contraction case
gives a strong mixing of the different layers, as does the high contraction case.
However, the dynamics behind the mixing seem to be very different, which is
obvious from the images of the jet.
Li et al. [132] studied stratified 3-layer headbox flow by an experimental
technique where salt was added to the centre layer and the conductivity pro-
file was measured across the jet thickness. In Figure 31 the left illustration
shows the principle, where the central layer has a higher conductivity. Due to
layer mixing there is a spreading (relaxation) of the conductivity profile
downstream. The overall mixing was summarised in a single mixing par-
ameter IMH. The right part of the illustration shows some conductivity pro-
files along the jet. They investigated the effect of slice opening, jet speed and
vane tip position on the degree of mixing. Mixing was reduced at higher slice
opening and flow velocity.
Li et al. [133] also studied the mixing phenomena in a 2-layer headbox,
using the same technique. In principle, the results obtained agreed with those
Figure 30 The effect of headbox contraction ratio (CR) on jet behaviour and layer
mixing. Scanned paper sheets (top) and images from the headbox jet (bottom).
Figure 31 Conductivity analysis of a jet from a 3-layer headbox (Left) and jet
conductivity profiles at 25 to 254 mm from slice opening (Right) [132].
for the 3-layer headbox. They also performed visualisations of the mixing in a
3-layered headbox using coloured water. After a comparison with the layer-
ing in the paper produced it was concluded that the main part of mixing takes
place in headbox and jet, and not during dewatering.
1. Vortex stretching, which is a result of the accelerating flow, will give vor-
tices aligned in the streamwise direction. This is a well-known mechanism
and the origin of vorticity can, for example, be the upstream conditions in
the nozzle or at the exit from the tube bank, see Hauptmann et al. [111].
2. Centrifugal instabilities (vortices) in the flow, which are caused by stream-
line curvature, see e.g., Matsson and Alfredsson [135]. These could be
so-called Dean or Görtler vortices. Robertson and Mason first proposed
the existence of centrifugal instabilities in the headbox flow.
3. Streaky structures, which grow inside shear flows due to hydrodynamic
instabilities could also occur at solid surfaces in the headbox, see e.g.,
Alfredsson and Matsubara [136].
Typically the width and/or height of streaks are of the order of the shear
layer thickness or the smallest size of the geometry perpendicular to the flow
direction. Hence, in wall-bounded flows the origin of streaks is usually closely
connected to the shear layer and thus has a size similar to its thickness. For
the case of a boundary layer flow the streaks typically will have a size com-
parable to the boundary layer thickness and for the case of fully developed
channel flows, i.e., the shear field covers the height of the channel, they will
have a size comparable to the channel height.
If the origin of the streaks is not coupled to shear fields, but instead invis-
cid mechanisms such as vortex stretching, the flow geometry itself will set the
size of the streaks. One example of this could be the flow in a nozzle (head-
box), where the final vortex diameter will be of the same size as the slice
thickness.
Söremark et al. [139] eliminated streaks on a paper machine by inserting a
small triangular bump on the bottom wall of the headbox contraction. This
small step would disrupt flow patterns (streaks) closest to the wall and due to
the small height of the bump it most probably altered the flow inside the
boundary layer. This method of streak prevention was earlier proposed by
Chuang [121] and can be seen in Figure 24 (centre), which shows a bump
mounted on the bottom nozzle wall.
Aidun and Kovacs [140] numerically simulated the flow of water inside a
headbox and identified the junctions between the converging nozzle walls and
the sidewalls as a source of secondary flow patterns, e.g., streaks.
A detailed experimental investigation of streaks generated by secondary
flow behind vanes was performed by Baker et al. [128]. They visualised the
flow patterns with Laser Induced Fluorescence and the images obtained
clearly showed MD streaks as well as CD structures/streaks.
In order to investigate the behaviour of the headbox jet at elevated speeds
(40 m/s) Schlupp et al. [141] performed visualisations of high-speed headbox
jets. These visualisations clearly showed streaks in the jet and the images
representing the jet surface were quantified regarding surface roughness
scales as a function of jet speed. This only showed a weak dependence of jet
speed. Experiments regarding streak generation by perforated plates, vanes
and slice-lip configuration were also performed; see Schlupp [142].
Lindqvist [143] made experiments concerning the creation of streaks using
water in a model headbox. The experiments were performed with Laser
Doppler Velocimetry and different visualisation methods. The visualisations
clearly showed the presence of large-scale structures in the flow, i.e., vortices,
which were generated by the tube bank. The measurements were performed
with water and showed the development of the boundary layer in the con-
verging headbox nozzle.
In order to understand how the flow from the headbox affects the jet it is
necessary to perform experiments as well as to make numerical simulations of
the jet flow. For the case of free jet flows, the simulations are a challenging
task since the location of the jet surfaces are not known, but are obtained as a
result of the simulation.
Li et al. [144] performed 2D simulations of the jet flow from a converging
nozzle.
In order to understand the behaviour of the streaks on the forming table
Aidun [145] modelled the flow from the slice and onto the forming table. This
included the flow of the jet. Also, the flow was quantified using high-speed
imaging and image processing methods, Aidun [146]. An example can be
found in Figure 33. The image is a spatio-temporal representation of the mix
on the wire. The vertical axis is the same as the CD and the horizontal axis
represents the time. The image clearly shows streaks in the mix surface.
In order to quantify flow visualisations of a headbox jet Söderberg [7] used
image processing methods, i.e., wavelets, to capture the behaviour of the jet
and mix on the wire. These visualisations were performed with transmitted
light as well as light reflected from the surface. The transmitted light visual-
isations were also used to obtain velocity distributions in CD/MD plane
using the so-called PIV technique. These showed streaks also in the velocity
distribution. Similar methods were used by Ono et al. [147] with the differ-
ence that they used light reflected from the suspension surface.
Söderberg and Alfredsson [138] have shown that streaks are not only ori-
ginating from inside the headbox nozzle, but can be created inside the jet
itself. This can be seen in Figure 34. The figure shows the flow of a plane
water jet at low Reynolds numbers visualised with the so-called shadowgraph
method. The velocity profile inside the jet gives rise to a hydrodynamic
instability (wave instability), which can be seen directly after the nozzle exit;
Figure 34a.
In Figure 34b, the velocity of the jet has been increased and the waves are
present only closest to the nozzle exit. After this short region of waves, the jet
experiences a breakup which partly causes a disintegration of the jet and also
gives rise to MD aligned structures. These have a CD spacing larger than the
waves and much larger than the jet thickness. The difference is one order of
magnitude.
Figure 34c shows a close-up of the break-up and the streak generation
inside the jet, which is visualised using reflective particles seeded into the
water. These particles are flat and thus they tend to orient with the shear in
the flow. Before the break-up the particles cannot be seen at all. This indicates
that there is no strong orienting shear before the break-up. After the break-up
clear white streaks can be seen, oriented in the MD direction and thus
implying an onset of strong shear, i.e., streaks.
Figure 34d shows a particle visualisation at the same jet thickness and
speed as in Figures 34bc. However, the nozzle is in this case formed as a slit
and the flow can be described as inviscid. The viscosity is then negligible and
boundary layers are not present.
The MD streaks seen in this image originate from within the nozzle and are
an effect of the vortex stretching mechanism, and the strong acceleration
tends to increase the strength of this vorticity and thus creates MD vortices.
This wave instability has also been shown to be present in full-scale head-
box flow, with water as well as with pulp suspensions; Söderberg [7]. In this
investigation, images taken of the jet were evaluated using a Morlet wavelet
transform; see Figure 35.
The transform is applied in MD. For one individual image (left) the wavelet
map is processed along 0.5 mm-wide MD-strips. The wavelet maps are then
averaged over all strips, which gives a spatial mean wavelet map. Further, in
this map all peaks along the wave number axis were recorded to form a peak
Figure 35 Left to right: Images at 5 m/s using reflected light, mean wavelet transform
and cumulative peak detection; a) water jet and b) mix jet with 0.35% softwood fibres.
event map. Finally, for both wavelet transform and peak event map averaging
was made from a sequence of 50 images.
The mean wavelet transform map (centre) gives an indication of the
dominant structures, while the cumulative peak detection map (right) shows
sub-dominant behaviour with high probability (could also be called peak
probability map).
The result, which can be seen in Figure 35, represents a jet velocity of 5m/s
and the top and bottom graphs are given by water and pulp suspension
respectively. The visualisation in this case is aimed at larger structures in the
flow; hence the image of water jet (top) does not show any presence of waves.
The top of the image is at a MD distance of 20 mm from the headbox and the
jet break-up can be seen. The white dots in the image are a result of this jet
break-up. If velocity is increased more dots will be visible.
The dots were, by visual inspection, found to be drops leaving the jet sur-
face as a consequence of the strength of the break-up. A ridge in the mean
wavelet map with a maximum at ∼4 mm can be clearly seen, which can be a
result of both the waves and the drops. It is also present in the cumulative
peak detection map.
The three graphs in Figure 35b show the flow of a fibre suspension at the
same parameters. In the image of the jet the fibre suspension (flocs) can be
clearly seen. Waves are present in this image, but are very hard to observe.
However, at higher velocity they could be clearly seen. The contour plot
shows a maximum at a wavelength of ∼20 mm, which is given by fibre floccu-
lation and in the peak map a maximum is found for wavelengths ∼4 mm,
which is a result of the surface waves.
acquired Ahlström, and their former became the “Valmet SymFormer MB”
unit.
The D-former was initially considered to generate a gradually thinning
space between the wires, through a gradually increasing force applied to the
bottom blades along the forming zone. This was supposed to make it possible
to gradually increase the dewatering pressure along the fabrics. In a presenta-
tion at the1989 Cambridge Symposium [1] it was explained for the first time
that instead of the gradually increased dewatering pressure suggested by
Dörries, a pulsating pressure was instead generated, by a zigzag movement of
the wires between top and bottom blades.
It was further suggested that the Dörries principle could have broader
applications:
The new principle has been applied to board making. It seems to have a
large potential also for high-speed papermaking. [1].
This basic principle of pulsating dewatering according to Figure 36 was
adapted in combination with roll dewatering in the STFI-Former. This was
initially presented at the PIRA Conference New technologies in multiply and
multilayer structures, one week before the start-up of the STFI-Former on the
EuroFEX pilot machine in June 1991 [151]. After some initial roll dewater-
ing, the fabrics were guided vertically upwards. Fixed blades were mounted
along the vertical tangent to the forming roll, against which the inner fabric
was run. On the opposite side, adjustable blades were mounted in a zigzag
manner; see Figure 37 [152].
In twin-wire forming, it is important to arrange the dewatering in a way
allowing automatic fabric position adjustment at changes in fabric separation
(stock quality or production changes) without significant changes in dewater-
ing pressure events. Therefore, the adjustable blades were mounted against
the movable outer fabric, and not against the inner fabric, the position of
which is initially defined by the forming roll surface.
One of the important development aims for the STFI-Former was three-
layer forming, and the idea was to form the outer layers during gentle
dewatering over the forming roll. This was followed by formation improving
pressure pulses that could break apart fibre flocs still remaining in the middle
layer.
It is important that all stages of dewatering are symmetrical, which is
the case when a suitable vacuum is applied at the roll surface and when the
blades are mounted symmetrically on both sides of the fabrics. The amount
of roll dewatering could be controlled by an adjustable cover angle along the
forming roll, and the individual blade pressure pulses could be controlled by
blade force adjustments.
One principal difference between pure roll dewatering and roll-blade
dewatering is the situation when the fabrics leave the roll. In roll-blade
dewatering, contrary to what is the case in pure roll forming, there is still mix
left between the two partially formed webs when the fabrics separate from the
roll surface, and the dewatering pressure disappears. This pressure drop will
cause an increase in mix velocity [153], see Figure 38.
At the same time, a critical table roll suction pulse will be generated when
the wires leave the roll surface. This may create a local velocity pulse as
indicated in Figure 38. If this pulse is strong, and the grammage of the web
on the inside of each wire is too low, sheet damage may arise. This mechanism
limits how small cover angle on the forming roll that can be used in practice.
It should be pointed out that in reality, the mix velocity during the roll
dewatering phase will not be generated by the long, constant amplitude pres-
sure pulse indicated in Figure 38, since the pressure build-up along the roll is
much more gradual; see further below.
Influence of gravity will mean that the mix velocity along the vertical
section at an average speed of 1200 m/min will decrease by nearly 60 m/min
for 2 m increase in height level.
Figure 38 Example of the variations in speed difference between mix and wires when
viscous effects are restricted to thin boundary layers. Broken lines indicate blade pulses.
Wire speed 1180 m/min, jet speed 1200 m/min, wire tension 6 kN/m, roll radius 0.8 m.
Voith developments
When the new Voith pilot machine in Heidenheim was started up in 1990
[154], it was set up for twin-wire blade forming, a technique which was pro-
moted by Voith at the time [155]. At the same time the development of the
STFI-Former was taking place, and thinking along the same lines, based on
the Dörries patent, also started at Voith. In a patent application from August
1989 [156], a roll-blade former is one of the alternative designs; see Figure 39.
However, in the patent there is no discussion of blade pressure pulses, and the
wires are described to gradually converge along the D-zone (marked II in
Figure 39).
Late in 1991, about one year after the start-up of the new pilot machine, a
rebuild was started to what eventually became the Douformer CFD unit
(“D” from Dörries) [157]; see Figure 40.
Besides the 45-degree angle against a vertical direction of the wire package
leaving the forming roll, there is one major difference from the STFI-Former.
Figure 39 Voith patent of roll blade former including adjustable blades [156].
Figure 40 Roll-blade Douformer CFD forming unit by Voith with loadable blades
against inner wire.
The fixed dewatering blades in the suction shoes are mounted against the
outer wire.
This means that the front-end position of the first shoe should be adjusted
depending on the wire separation distance, leaving the roll, if the same initial
pressure pulse should be obtained. However, the shoe position is not adjusted,
once the suction shoe has initially been installed according to instructions.
Application of the Duoformer CFD to printing papers indicated improved
formation and retention in comparison to previous designs [158,159].
The Duoformer CFD was later modified to DuoFormer TQ (Total
Quality). A main objective was to increase web dryness entering the press
section, which was obtained by increasing the wrap angle of the top (outer)
wire around the couch roll, and by the introduction of an extra suction box
after the couch roll [160,161]. A further development is the DuoFormer TQv,
where “v” indicates a vertical wire run, leaving the forming roll [162,163].
Early 2001, Moser [164] described the main Voith twin-wire forming
designs, together with a discussion of important paper properties. A total
number of 190 DouFormer D units were reported to have been delivered at
the time.
Voith also suggested a simplified version of roll-blade forming by position-
ing some adjustable blades against the outer wire on a roll former; see Figure
41. This design was tested on an industrial roll former producing newsprint.
A clear improvement in large-scale formation was obtained, like that with
blades placed after the roll.
However, this concept breaks a fundamental rule of twin wire forming:
“Never physically control the positions of both wires at any location along
the dewatering section”. What happened with time was a gradual increase of
grammage streaks.
This was caused by the slight friction wear of the blade surfaces, which in
turn generated local changes in roll-to-blade distance across the CD and thus
also changes in local wire separation across the machine, causing grammage
streaks. When a blade is applied against one of the wires, the opposite wire
should be able to automatically adjust its position to compensate local blade
surface deviations.
Figure 41 Adjustable blades mounted against forming roll outer wire. Voith
patent [165].
Valmet developments
Valmet considered the application of the Sym-Former MB unit with adjust-
able blades, after some initial roll forming, and a patent was applied for in
May 1989 [166]; see Figure 42.
The blade section was designed according to the Alform MB principle for
Fourdrinier machine rebuilds, which Valmet had inherited in the acquirement
of Ahlström. In the patent there is, however, no disclosure of pressure pulses
generated by the adjustable blades; wire convergence in the zone between the
stationary and the adjustable blades is described as “mainly linear”.
Figure 42 Roll-blade former according to Valmet patent [166]. Item “127” includes
the loadable blades against inner wire. This is basically the SpeedFormer MB design.
Figure 43 Forming roll followed by suction shoe against the inner wire and loadable
blades against the outer wire in the Valmet Optiformer [168].
former case were acting against the inner wire. Thus the fixed blades, like in
the STFI-Former (Fig 37), are now acting against the inner wire, the entering
position of which is always defined by the roll periphery, independently of the
separation between the two wires. Therefore no adjustment of the position of
the first fixed blade is needed, when running conditions are changed. This is
most probably the main reason why better formation is obtained on the
OptiFormer.
Beloit developments
The traditional design of the Beloit Bel Baie machines included initial
dewatering over a forming shoe consisting of nine blades. In 1995 Bel Baie IV
design was presented, in which the initial dewatering takes place over three
wide (ca 150 mm) blades with large radius of curvature [169]. This introduces
a gentler initial dewatering in comparison with the sharp pulses over the
traditionally narrow (ca 15 mm) Bel Baie blades. It can be seen as a replace-
ment of initial roll forming, with the drawback of only one-sided dewatering
as well as considerable friction against the contacting wire.
A top wire unit for rebuild of fourdrinier printing paper machines, consist-
ing of curved, inverted suction boxes on top and pneumatically loadable
blades below the wires was described in 1995 [170].
6 WEB FORMING
6.1 Random sheets
The sheet formation number FRandom (coefficient of variation) for a sample
with identical, randomly distributed fibres is determined by the square root
Figure 44 Top-ply roll-blade board forming design; the Voith DuoformerTop [177].
of the inverse mean number of fibres at a point in the plane of a paper sheet;
see Equation 5:
冪
wFibre
FRandom = (5)
wSheet
Figure 46 Tensile strength index for single- and double-layered sheets. Laboratory
sheets of long fibre pulp; increased forming concentration above arrow [192].
ΔP
= av + bv2 (7)
L
Danby et al. made several studies of the influence of the forming fabric on
print quality. Initially, the different geometries of single, double and triple
layer fabrics on print evenness were demonstrated [204]. Periodic marking in
split sheet surfaces was evaluated. It correlated highly with respective fabric
surfaces and print marking [205,206]. With consideration of the fibre length
distribution, a correct choice of fabric mesh pattern (opening size and orien-
tation) had a large effect on retention level. A computer model was also
developed to predict print quality from the image of the surface layer of
paper samples after sheet splitting [207].
A similar flow loop as that of Herzig and Johnson was used by Jong et al.
[208]. The specific flow resistance of sheets of different grammage, formed by
LWC pulp, was measured. It could be noted among their results that in the
case “Closed loop Screen B”, a very high flow resistance was recorded at
20-30 g/m2, which dropped drastically at further fibre addition; see Figure 50.
This might indicate “sheet sealing”.
If no flow resistance were imposed by further fibre addition, the resistance
curve would follow the added broken curve. In reality, the extra fibres do
contribute to sheet resistance, but only to a degree corresponding to the
distance from the broken curve.
Figure 50 Effect of basis weight BW on mat flow resistance SFR for a LWC pulp
(Figure 7 in [208]). Added broken curve follows the equation SFR = k/BW.
Fabric design may also be very important regarding rewetting. When the
actual web build-up is finalised, web dryness is increased over vacuum boxes
and couch roll. If water after this still remains in the fabric structure, a
substantial part of this water is likely to be sucked back into the paper web,
before and/or after its transfer to the pickup felt. Web rewetting by a single
layer forming fabric was studied by McDonald [209] on a pilot machine.
Samples for dryness evaluation were collected after the couch, in a grammage
range of 25–88 g/m2 using newsprint furnish. Rewetting was evaluated using
the Sweet method and gave a rewet value of 55 g/m2. This is a significant
amount of rewetting, especially for a low grammage product.
Simulation of initial fibre retention by the forming fabric has been made
using software developed to model the forming fabric in 3D, given the yarn
and weave parameters [210,211]. The geometric probability of initial reten-
tion for a fibre is calculated based on its given length and orientation.
insensitive to degree of beating and beater type. A model was presented for
the build-up of a fibre mat from a fibre suspension based on the response
determined in the above drainage device [217].
Mantar et al. [218] used a constant pressure drainage tester to study
dewatering of fibre suspensions of chemical and mechanical pulps using 17
kPa dewatering pressure. Dewatering rate was evaluated by measurement of
the drained volume. Influence of grammage (up to ca 400 g/m2), suspension
concentration (0.1–0.9%) and fines was included.
Ramarao and Kumar [219] modelled gravity drainage of pulp suspensions.
During gravity drainage, pressure, concentration and filtration resistance will
pass through a maximum with respect to time.
The model can be used to study drainage in conventional laboratory
devices such as hand sheet formers and freeness testers.
Wildfong et al. [220] studied filtration mechanisms of sheet forming. Con-
stant pressure filtration was generated by opening a valve to a vacuum cham-
ber with a pressure of −6 kPa. Dewatering rate was evaluated using optical
triangulation of the slurry surface. Realistic suspensions were taken as head-
box samples from newsprint, LWC and fine paper operations, but since the
retention values obtained were as high as 80–90%, fines content in the webs
formed were over-represented. Industrial like conditions (slice opening) were
used for initial slurry height. Although considerably more realistic than other
dewatering experiments reported, a slow initial dewatering (acceleration from
standstill) may be the most important deviation from industrial conditions
with a sudden jet to fabric impact. As related above, the initial dewatering
velocity may have an important influence on the build-up of the first fibre
Figure 52 Viscous resistance coefficient values a for different fibre suspensions [220].
7 DEWATERING PROCESSES
7.1 Jet impingement
The jet impingement on a forming fabric is a critical part of the paper forming
process, and has been studied experimentally and by modelling.
Shands [122] studied mix jump at jet impingement using linear wave theory,
and experimentally using laboratory flow models and pilot paper machine
trials. The experiments confirmed the theoretically suggested mechanism that
mix jump can result from wave amplification caused by the acceleration
changes that are produced as the mix jet lands on the wire. The effects of
headbox generated turbulence, free jet length, jet speed and curvature were
studied experimentally (see Figure 25).
A forming board arrangement with a radiused lead blade that can be used
to control mix jump at high machine speeds was also described.
Audenis [224] modelled the impingement of a plane inviscid liquid
jet on a forming fabric at different jet incidence angles. The deformability
and the porosity were considered separately. The present method extends
previous works in that part of the fluid is allowed to flow out of the
computational domain along the fabric and a more realistic boundary
condition has been applied at the porous fabric in terms of a permeability
law.
Jong [214] made a numerical and an experimental analysis of jet impinge-
ment on one wire and in a twin-wire nip. Measurements were made on the
Paprican pilot machine. There was a free wedge zone against one wire, also in
the twin wire case before contact with the forming roll took place. Reasonable
results were obtained, with roughly twice the amount of dewatering evaluated
numerically in comparison with experiments.
Dalpke, Green and Kerekes [225] modelled the impact of a jet onto a flat
wire using CFD software. Their main aim was to study the entrance region of
a twin-wire roll former, but the initial investigation treated the landing of a jet
on a flat wire. One viscous and one inertial term were used to characterise
wire flow resistance. In the main part of the calculations, fibre mat resistance
was neglected. Jet angle and velocity were important parameters. It was
found that the main part of dewatering took place within a distance equal to
the jet thickness, and that the pressure event further downstream was little
influenced by the different variables settings.
450 m/min a jet excess speed of 7% was reduced to zero within 7 m (ca 1 s). It
was also possible to evaluate MD and CD velocity components at jet exit.
CD velocities agreed with fibre orientation misalignment angles of the
paper produced. Loewen [227] also discussed fibre orientation anisotropy and
misalignment in relation to jet-to-wire speed difference and headbox edge
flow adjustments.
Farnood et al. [228] used a laser velocity sensor to evaluate MD mix speed
and table activity and a gamma radiation gauge to evaluate drainage profile.
They studied effects of slip velocity, mix concentration and table activity on
paper formation. They suggested that such equipment could be applied for
on-line control of formation.
Kiviranta [229] made several studies of table activity in the Fourdrinier
section, and its effect on paper formation. Initially a photo clinometric
method to quantify mix surface irregularity was developed [230]. It was based
on a low angle sideways incident stroboscope flash, and a recording CCD
camera mounted over the wire section. Three main parameters were found to
be important in characterising activity: surface roughness, correlation length
and scale.
The factors affecting stock surface structure and paper formation for the
middle ply of folding boxboard was studied on a pilot machine [231]. The
same pilot machine was used to study the effects of mix concentration and
gravity foil parameters on fine paper formation [232] and on linerboard for-
mation [233]. These results were later also summarized [234].
The new method was tested on an industrial fine paper machine [235]. It was
found that table activity should be kept low in the initial dewatering stages. A
too severe activity decreased retention, which in turn automatically increased
retention aid addition, generating a higher degree of fibre flocculation.
Kiviranta and Dodson [53] used the pilot machine data by Kiviranta
described above, and included the Crowding factor N to replace the fibre
concentration. This improved the possibilities to predict formation levels. An
increase of N by 10 units corresponded to roughly 10% increase of local
grammage variations, which also corresponded to the range of influence on
formation by change of foils angles. The regression equations developed
could describe the effect of the parameters considered within a rather wide
range of operating conditions.
Foulger [236] investigated the submerged drainage concept of dewatering.
The new submerged drainage box replaces the conventional low vacuum box.
The main difference is that air is avoided by the use of a water filled box under
gentle vacuum. An increased dewatering capacity in the low mix concentra-
tion range, 1–2%, was reported, combined with a significant increase in the
retention level.
Since the actual inward flow was much higher than simulated, they concluded
that the inner web resistance was low. This could be a result of washing out
effects by the blades, similar to that on a fourdrinier machine.
Bando et al. [39] also studied the effect of pressure pulses on orientation
anisotropy.
At increased jet to wire speed difference, they found that changes in
anisotropy only took place in the middle half of the sheet. They calculated
local flow velocity between the wires (see Figure 55), and evaluated cumulative
shear factor (m2/s) by integrating the absolute difference between wire and mix
speeds along the forming section. Fibre orientation anisotropy in the middle
layer increased with cumulative shear factor, but the effect of high amplitude
pulses was relatively much higher than that of lower pulses.
Figure 55 Velocity profile of pulp slurry between top and bottom wires as a function
of distance from jet impingement position for 5, 9 and 18 blades respectively [39].
The shape of blade pressure pulses was also studied by Zhao and Kerekes
[245]. A mathematical model was developed, assuming inviscid flow, a thin
blade, zero wire bending stiffness and constant permeability; see Figure 56
(left). They also made experiments on the Paprican pilot machine, and meas-
ured the pressure events using tappings along a flat blade, with wire deflection
over the back end; see Figure 56 (right).
Zhao and Kerekes [246] also tried to model the formation improvement
generated by blade pulses. They compared formation values (NUI) in roll
and roll-blade forming using a range of blade pulse amplitudes. They finally
correlated the formation improvement NUI with the calculated total mix
displacement due to the blade pulses and claimed a positive correlation.
Figure 56 Left: Calculated pressure pulses at different wire speeds along a flat blade
with wire deflection over downstream end. Right: Comparison between calculated
and measured pressure pulses [245].
However, the spread in the data was much too high to motivate such conclu-
sions. It would be useful to characterise also formation scale, if conclusions
about the sheet forming mechanisms are to be drawn.
Nigam and Bark [247] considered the two-dimensional flow field between
two wires deflected over a curved blade. The case with impermeable wires and
irrotational flow was analytically solved; see an example in Figure 57.
Zahrai et al. [248] studied different aspects of twin-wire forming. Initially
a two-dimensional model was applied using a thin blade, and effects of
permeability and wire bending stiffness were considered [249,250]. Local suc-
tion pulses downstream of a blade were predicted. Analytical calculations
were compared with experimental results from the EuroFEX machine using
different blade shapes [251]; see Figure 58.
Green [252] reinvented the fact that the blade force (integrated blade
pressure) is proportional to the outer wire deflection angle over a blade.
Green et al. modelled different aspects of blade related pressure events and
studied:
• two-dimensional simulation of pressure pulses in blade formers [253], (like
Zahrai, see above) they found a downstream vacuum pulse;
• influence of wire stiffness, inner and outer wire tensions and blade width
[254];
• numeric analysis of blade pulses, in comparison with earlier analytical
Figure 57 Streamlines for the velocity field between two wires deflecting over a
curved blade (tip length 50 mm). No deflection of inner wire on ingoing and outgoing
side. Scales in m [247].
Figure 59 Modelled blade pressure peaks (left) and local mix velocities (right) in a
blade arrangement with 5 initial blades followed by 4 loadable blades and 5 stationary
blades. Wire tensions 6 kN/m and blade loads 0.5 kN/m [258].
Suction shoes
It was suggested by Odell, Pakarinen and Luontama [259] that when a twin
wire sandwich passes over a suction shoe, the wires will be bent into the shoe
openings by the suction forces; see Figure 60. The hypothetical result would
be an increased outer wire deflection around blade back and front edges,
which in turn would create increased pressure pulse amplitudes between the
wires.
Green [260,261] made an effort to model the mechanisms of suction shoe
dewatering, starting from a wire geometry as shown in Figure 61.
Roshanzamir, Green and Kerekes [262] attempted to model the suction
shoe event with a two-dimensional simulation, based on the 1-D method
earlier tried by Green. The argument that good agreement with the 1-D
Figure 61 Hypothetical paths for two wires travelling along a suction shoe opening
[261].
method lends credence also to the 2-D method does not seem convincing.
Also in this case a considerable suction is predicted inside the outer web. The
modelled geometry of the wire runs in the different cases treated is, however,
not reported.
However, the physical model behind Figures 60 and 61 is incorrect since the
force bringing down the outer wire into the opening is negligible. The wire
will, therefore, travel along a mainly straight path between two adjacent
suction shoe blades.
When there is still free mix between the inner and outer webs, the web on
the inner wire will absorb the main part of the suction pressure, and the inner
wire would bend. The free mix would then support a vacuum high enough
to loosen the web from the outer wire. At a later stage, there will be one solid
web between the wires, which would loosen from the outer wire due to an
inside vacuum.
Suction application in the blade section of a twin-wire former will have a
major effect, previously not reported, on the generation of blade pressure
pulses.
In a blade arrangement with fixed blades on one side and loadable blades
on the opposite side, a pressure pulses is created in front of a loaded blade.
(Also about half the effect will occur at the two surrounding blades on the
opposite side, but this will not be further discussed now.) The basic reasons
for this mechanism is that when a loadable blade moves the contacting wire
closer to the opposing wire, the space for mix between the wires is reduced. To
still maintain the same rate of mix flow, a high-pressure zone will develop
upstream of the blade, and that pressure zone will locally move away the
opposite wire, so that the cross sectional area is mainly restored.
If now suction is applied beyond the opposite wire, the low pressure will
help to locally displace the opposite wire, and thus increase the cross sectional
area for mix flow. At some suction pressure level, the displacement of the
opposite wire will be sufficient to let the mix pass without the help of the
blade generated pressure pulse between the wires mentioned above. The blade
pressure pulse will thus disappear. The applied blade load will instead be
completely balanced by the local deflection of the contacting wire, which will
then increase in comparison with the deflection without suction (and then
increase the pressure pulses at the two surrounding blades on the opposite
side).
The pressure event during the initial part of roll forming has not yet been
modelled theoretically, and the local pressure minimum just before the final
maximum is also so far unexplained (see further below).
Martinez developed a physical model to predict the dewatering through the
two wires in twin-wire roll forming. A series of force and mass balances were
applied to the web, and expressions were derived for web thickness and solid-
ity based on fibre and process conditions. Using the measured pressure event,
dewatering rates were estimated with Darcy’s law. A series of experiments on
the EuroFEX machines were performed, and one test was used to calibrate
the parameters. The numerical solution predicted flow rates within 10%; see
Figure 63.
Boxer applied modelling techniques to a roll-blade former. He modified the
method by Martinez for roll dewatering, by replacing the numerical solution
method by an analytical one [264]. It is claimed that with the new method,
knowledge of the actual pressure event during dewatering is not required for
the solving of the dewatering flow rates [265].
Zahrai, Martinez and Dahlkild [266] presented a physical model to esti-
mate the thickness of the web in twin-wire formers with nearly constant
drainage pressure. Web thickness was found to be proportional to the square
root of forming distance. The constant of proportionality was related to
parameters such as compressibility and permeability of the web, headbox
consistency and machine speed.
Figure 63 Comparison of the estimated to the predicted dewatering rates from both
the inner (A) and outer (B) fabric [263].
Zahrai, Bark and Martinez [267] modelled the dewatering from a curved
converging channel with moving walls, of which the outer one was permeable.
A numerical solution was obtained with increasing wall excess speeds relative
to the suspension for two fibre suspension models: one power law version and
one more “fluid-like”. It was shown that with increasing speed difference, the
fluid-like model resulted in an increased fibre concentration of the web at the
interface to the suspension, which was not the case using the power law
model.
Turnbull et al. [268] developed a 1-D model for a Crescent former (roll
former with one-sided dewatering) to study dewatering events, and specific-
ally the influence of longitudinal disturbances. Despite heavy damping,
the model showed significant amplification of disturbances near resonant
frequency. The model predicts that 2% disturbance in jet velocity then
may result in nearly 7% variation in basis weight. The wavelength for such
variations approximately equals the forming roll radius.
Chen et al. [269] modelled the CD mat thickness disturbances on a Cres-
cent former using a two-dimensional dynamic model with viscosity. The
study examined the CD variations due to the influence of major classes of
steady disturbances, and the coupled MD and CD variations due to dynamic
process disturbances in the fluid jet. Among many results, those from
unsteady disturbances show that the coupled MD and CD variations are
magnified near the fundamental natural frequency of a translating wire for
long wavelength disturbances.
Jong [214] made some roll forming modelling attempts and compared
the results with measurements on the Paprican pilot machine. One main
conclusion was that the model should be further developed before good pre-
dictions of dewatering could be expected.
In a recent overview of twin-wire forming, Malashenko [270] discussed a
typical roll pressure event in a roll-blade former and commented “Prior to
exit, there is an ever present and unexplained reversal in pressure . . .”. It
could be pointed out that such a local minimum is also present at the end of
the B-section in Figure 62.
Malashenko also stressed the importance of relative motion between wires
and connected web shear to different wire tensions and running radius over
rolls; see Figure 64.
Figure 64 Differential fabric displacement and web shear in twin-wire “S” path, [270].
Wildfong et al. [271] modelled drainage during roll forming, and validated
results using pilot paper machine data. Stock drainage data were evaluated
according to methods described above [220,221]. Roll dewatering pres-
sure was assumed constant, and the vacuum in the forming roll was neg-
lected. Prediction accuracy of drainage flow was within 15% of experimental
values.
A pilot plant study of roll dewatering of reslushed newsprint was made by
Gooding et al. [272]. They found that dewatering comprised two main parts,
one momentum driven and one tension-driven. Dewatering rates were
obtained by scooping with a special collection device, and further by flow
balances based on wire position estimations based on a probe recording
mechanical bending of the probe assembly. No separate measurement was
made of dewatering to the roll side, and the concluded values resulted in
some peculiarities. Normal fabric tensions were applied, but since the roll
diameter was only about half that of a modern industrial installation, the
dewatering pressures were twice the normal levels. Recorded retention levels
were high; for a fine fabric over 85% was obtained. Pressure events were
recorded using an air purged capillary tube introduced through the headbox,
and stopped at different positions along the forming zone, the same tech-
nique as that used by Beloit and Mitsubishi [242]. Pressure recordings reveal
some uncertainty about the probe tip location; sometimes it seems to have
been located inside the forming web.
C = b + m tanh(ct) (8)
Figure 65 Web solids as a function of dwell time for flat box at three vacuum levels.
The lines are drawn using Equation 8 [273].
middle curve for the relevant time period, and then move left to the third
curve, etc.
By application of this principle, he demonstrated considerable potential in
reductions of drag forces (fabric wear) and vacuum energy consumption
through improved strategies of vacuum application, replacing traditional
settings.
Neun and Fielding [274] studied vacuum box dewatering by field meas-
urements. They concluded that corresponding laboratory results were rea-
sonable absolute predictors. Increased grammage had an effect on slowing
down dewatering rather than on achievable dryness. Furnish had a large
effect, and higher temperature improved dryness.
Neun also evaluated the parameters m and c (Equation 8) for newsprint,
OCC and Kraft furnish at different vacuum levels [275,276,277].
Räisänen [278,279] studied flat box vacuum dewatering using the “Moving
Belt Drainage Tester”, MBDT (see description below). Using this equipment,
the same type of vacuum pulses are applied both during forming and vacuum
dewatering.
The process was considered to be basically a wet pressing event, and the
solid content C as a function of time t was described based on a first order
viscoelastic model
冢 冣
t
C = C0 + b 1 − e− r + dt (9)
The linear term was added to account for the effect on water removal of the
through-flow of air. The effect of different furnish parameters were studied
[280] and the laboratory evaluations were successfully compared with the
results for a pilot [281] and a commercial [282] paper machine respectively.
If, in fact, wet pressing is the dominating effect for water removal on
vacuum boxes, this part of the process should preferably be treated together
with the press section. Traditionally this has however not been the case, and
therefore this part is here still included in the forming section.
The MBDT was later also used by Mitchell and Johnston [283] to study
vacuum dewatering.
Shands and Hardwick [284] studied flat box vacuum dewatering on a pilot
machine for newsprint at commercial speeds. Dryness increase was found to
be mainly proportional to “vacuum impulse”, i.e., vacuum level multiplied
by dwell time. Also airflow requirements and drive loads were found to be
proportional to vacuum impulse.
Jones [285] developed a dynamic simulation model for flat box vacuum
dewatering using the experimental results by Neun (mentioned above). The
model is based on flow through compressible porous media. What appear to
explain the limitation of the flat box are the forces required to remove water
from the smaller pores of the sheet. Dewatering stops at the point at which
the surface tension forces in the pores exceed the applied vacuum. This can
explain why solids content levels off regardless of dwell time but increases
with increasing vacuum and temperature. Increased grammage slows drain-
age but does not reduce maximum solids content.
8 LABORATORY FORMING
For the channel flow on top of the wire, the Froude number Fr has signifi-
cance, with
v2
Fr = (10)
gd
Sivén and Manner [295] designed a high-speed retention tester, the HSR-
Tester, based on a moving headbox and stationary twin-wire arrangement.
Inside the inner wire, drainage foils are mounted on a radius defining the
inner wire radius, and the set-up is rotated at ca 5 r/s. Outside of the inner
wire, the likewise stationary outer wire is tensioned. In between, the headbox
is placed, and when a sheet forming event is started it is retracted at ca 90 m/
min along a half-circle formed track, leaving the mix mainly stationary to be
drained through the two wires with the foil pulses agitation from inside. The
quality of the paper regarding formation and mechanical properties are not
reported so far. Its main aim is to study retention as a function of retention
aids addition [296].
Hammock and Garnier have developed a laboratory apparatus simulating
industrial twin-wire forming [297]; see Figure 69. The headbox width is 160
mm. The forming roll (diameter not specified) surface has 2 × 2 mm2 grooves
separated by 1 mm-wide ridges, to allow also inwards dewatering. After the
roll fixed blades on both sides and vacuum boxes bring final sheet dryness up
to 8–10%, which allows sheet sampling. The mix is pumped from a 400 L
reservoir, and bypassed back to the reservoir via a three-way valve before and
after the trial. Trial time is within the range 5–50 s. In a typical trial slice
opening was 4 mm, wire speed 320 m/min and mix consumption 140 L.
The EuroFEX plant is today the only laboratory plant in which indus-
trial running conditions can be simulated if chemical additions and/or
multi-ply forming are involved, due to continuous fresh stock addition as
well as continuous recovery of material from excess white water. The dif-
ference from a pilot plant without these two design features was recently
demonstrated by Nalco Chemical Co [302]; see Figure 70. They made
similar test runs in the EuroFEX and the Beloit pilot plants. It is clear
from these tests that in the EuroFEX plant the white water chemistry
levels out between the different test points and that the levels recorded are
higher.
Figure 70 “Comparison of the SLM/FBRM mean chord length traces obtained for
XPM3 (Beloit) and EuroFEX pilot paper machines. The marks on the figure indicate
changes in retention chemistry or dose.” Figure 12 from [302].
• The most obvious way of improving both formation and paper strength
is to reduce forming concentration, assuming that current drainage
limitations are not superseded.
• An often-used method to worsen formation in hand sheet forming is to
impose a delay time before drainage. Formation as well as strength will
then deteriorate.
On the contrary, there are several occasions where the opposite situation
will be true, i.e., formation improvements will not be accompanied by
strength improvements:
• Excessive addition of retention aid can improve formation, but paper
strength will be unchanged.
• Manipulation of the activity on a fourdrinier wire may improve formation,
but this will normally be associated with a reduction in paper strength.
• An industrial method to improve formation in twin-wire forming is to
introduce blade pressure pulses. However, in this case paper strength will
deteriorate [117].
A “general rule” was formulated by Fredlund [303] as a summary to a
range of EuroFEX pilot plant experiments:
Strength potential in forming lies in the quality of the headbox jet. Manipula-
tions at later stages can improve formation but not paper strength.
different forming principles and jet-to-wire speed ratios; see Figure 74. They
motivated the different profile shapes by different dewatering rates and differ-
ent speed differences between web and mix:
• the low degree of anisotropy towards the wire surfaces is due to the fast
drainage in those areas, which leaves little time for fibres to orient due to
shear;
• the lower degree of anisotropy at the centre with two-sided dewatering is
due to reduced shear velocities due to viscous damping;
Figure 74 Layered orientation anisotropy during “rush” (left) and drag (right)
forming conditions for three forming principles [43].
1. Anisotropy;
the definition of “anisotropy” in this work, Equation 11, differs from that
in Equation 2; see Figure 8.
Figure 76 Layered orientation structure for roll-blade forming at 1200 m/min, LWC
50 g/m2. “Anisotropy” (Left) and Standard deviation of “anisotropy” (Right). Jet/wire
ratio on the low side; 1.02 corresponds to minimum shear during dewatering [315].
Figure 77 Layered floc index (left) and misalignment angle (right), three different
thickness ranges from sheet bottom to top; different jet/wire ratios, data as in
Figure 76 [315].
web centre. This could explain why shear during dewatering mainly has a
positive effect in the late dewatering phases – this is simply when the flocs are
to be dewatered.
The results reported by Erkkilä et al. also include rush conditions, but
space does not allow its presentation at this time.
The effect of differences in vacuum levels in forming roll surface and form-
ing shoes was also studied, but had only limited influence on sheet structure.
The mounting of vanes inside a headbox nozzle is well known to have a
decreasing effect on fibre orientation anisotropy in the final paper product.
As shown above in Figure 22, vane insertion reduces fibre orientation
anisotropy in the headbox jet.
Erkkilä et al. [315] studied the effects on the paper of headbox vanes and
roll-bade forming; see Figure 78.
It is obvious that there is no influence of vane length itself (with the actual
dimensions), only of end shape. However, the turbulent eddies generated by
the blunt vane end probably lasted all the way to the dewatering stage. The
reducing effect of turbulence on anisotropy and the negative effect on floc
index are clearly demonstrated.
Figure 78 The effect of different vanes on layered “anisotropy” (Left) and floc index
(Right). The 500 and 600 mm vanes end 220 and 120 mm respectively, before slice
opening [315].
Figure 79 Fibre orientation map for layer next to sheet top using blunt tip vanes;
data as in Figure 78 [315].
forming on a fourdrinier machine. Page and Hergert also discuss the advan-
tages for paper mechanical properties using different layering in alternative
cases.
During the 1990s several installations of 2-layer headboxes on linerboard
machines have been made. This makes it possible to manufacture white-top
liner with only one forming unit.
During the last decade much work has been performed to improve the
quality of 3-layered printing papers (see section Headboxes). One main prob-
lem in comparison with 2-layer linerboard forming is the low grammage of
the surface layers; brighter surface layers cannot optically cover a darker
centre layer. The centre layer brightness therefore should not deviate too
much from that of the surfaces.
Häggblom-Ahnger [320] made several pilot-plant trials to study three-layer
forming of printing papers. Häggblom-Ahnger and Eklund [321] studied the
effects of a CTMP middle layer in a 3-layer copy paper. Positive effects on
mechanical properties were found, but the surface layer cover was not even
enough.
In a study of the location of softwood fibres (30%) in a three-layer office
paper bending stiffness was found to increase with a central placement;
ACKNOWLEDGEMENTS
REFERENCES
27. Daunais, R. and Garner, R., “The NUI formation tester – an evaluation, modifi-
cation and comparison with other techniques”, in Inernational Paper Physics
Conference, Mont Gabriel, p. 43 (1987).
28. Kallmes, O. and Ayer, J., “Light scanning system provides qualitative formation
measurement”, in International Paper Physics Conference, Mont Gabriel, p. 209
(1987).
29. Popil, R. and Jordan, B., “The wavelength dependence of transmitted light forma-
tion measurement”, in International Paper Physics Conference, pp. 61–69 (1995).
30. Lloyd, M. and Norman, B., “Effect of headbox slice geometry during the stratified
forming of woodfree paper”, Nord. Pulp Paper Res. J., 12(4): pp. 252–259 (1997).
31. Moineau, D., Bloch, J-F. and Eymin, G., “On-line 2D formation sensor: A new
tool to evaluate the quality of paper formation”, in International Paper Physics
Seminar, Grenoble, pp. 29–38 (2000).
32. Hiertner, M., “Basis weight varies faster than on-line systems can measure”, in
Control Systems ’98, Porvoo, p. 307 (1998).
33. Ferguson, H., “Full sheet image system becomes control reality”, Pulp & Paper,
71(10): p. 75 (1997).
34. Mathias, T., “Paper industry’s first full sheet imaging system”, in 52nd Appita
Annual Conference, pp. 429–433 (1998).
35. Hagart-Alexander, C., “Wet end measurement system for MD and CD quality
monitoring and control”, in 86th Annual Meeting, PAPTAC, Monteral, pp.
231–233 (2000).
36. Niskanan, K., Kajanto, I. and Pakarinen, P., “Fibre orientation”, Paper Physics,
Papermaking Science and Technology, Fapet OY, Helsinki, 16(1.4): pp. 37–50
(1998).
37. Titus, M., “Ultrasonic technology – measurements of paper’s orientation and
elastic properties”, Tappi J., 77(1): pp. 127–130 (1994).
38. Nomura, T., “Introduction of new SST-3000 and development of sonic sheet
tester uses”. Japan Tappi J., 48(1): pp. 215–219 (1994).
39. Bando, T., et al., “Drainage mechanism on a twin-wire former. Part 3: Depend-
ence of the structural anisotropy of paper sheet on the pulsating drainage”, Japan
Tappi, 49(49): pp. 841–848 (1995).
40. Masuda, K., Hasuike, M. and Bando T., “Characterization on 3-D geometric
arrangement of fibres in a paper sheet”, in ISF, Yokohama, p. 243 (1994).
41. Hasuike, M., Kawasaki, T. and Murakami, K., “Evaluation method of 3-D
geometric structure of paper sheet”, J. Pulp Pap. Sci., 18(3): pp. J114–120 (1992).
42. Hasuike, M., Masuda, K and Bando, T., “Characterization of three-dimensional
sheet structure”, Mitsubishi Heavy Industries, Technical Review, 32(3), 127–131
(1995).
43. Erkkilä, A.-L., Pakarinen, P. and Odell, M., “Sheet forming studies using layered
orientation analysis”, Pulp Paper Canada, 99(1): pp. 81–85 (1998).
44. Jansson, M., “Fibre orientation anisotropy – variations in the z-direction”, (in
Swedish). Pulp and Paper Chemistry and Technology, KTH, Stockholm, MSc
Thesis (1998).
45. Lloyd, M. and Chalmers, I., “Use of image orientation analysis technique to
investigate sheet structural problems during forming”, in APPITA, Melbourne,
pp. 495–502 (2000).
46. Scharcanski, J. and Dodson, C., “Local spatial anisotropy and its variability”, J.
Pulp Pap. Sci., 25(11): pp. 393–397 (1999).
47. Thorpe, J., “Exploring fiber orientation within copy paper”, in International
Paper Physics Conference, San Diego, pp. 447–457 (1999).
48. Xu, L., Parker, I. and Filonenko, Y., “A new technique for determining fibre
orientation distribution through paper”, in International Paper Physics Confer-
ence, San Diego, pp. 421–427 (1999).
49. Kao, S. and Mason, S., “Dispersion of particles by shear”, Nature, 253(February
20): pp. 619–621 (1975).
50. Norman, B. et al., “Hydrodynamics of papermaking fibres in water suspension”,
in Fibre-Water Interactions in Paper-Making, Oxford, pp. 195–250 (1977).
51. Steen, M., “Modeling fiber flocculation in turbulent flow: a numerical study”,
Tappi J., 74(9): pp. 175–181 (1991).
52. Kerekes, R. and Schell, C., “Characterization of Fibre Flocculation Regimes by a
Crowding Factor”, J. Pulp Pap. Sci., 18(1): pp. J32–38 (1992).
53. Kiviranta, A. and Dodson, C., “Evaluating fourdrinier formation performance”,
J. Pulp Pap. Sci., 21(11): pp. J379–384 (1995).
54. Kerekes, R. “Perspective on fibre flocculation in papermaking”, in International
Paper Physics Conference, Niagara, pp. 23–31 (1995).
55. Andersson, S., “Netwok disruption and turbulence in fibre suspensions”, Dep of
Chemical Engineering and Design, CTH, Gothenburg, PhD Thesis (1998).
56. Wikström, T., “Flow of pulp suspensions at low shear rates”, Dep of Chemical
Engineering Design, CTH, Gothenburg, Licentiate Thesis (1998).
57. Wikström, T. and R. A, “Yield stress of pulp suspensions”, Nord. Pulp Paper Res.
J., 13(3): pp. 243–250 (1998).
58. Andersson, S. and Rasmuson, A., “The network strength of non-flocculated fibre
suspensions”, Nord. Pulp Paper Res. J., 14(1): pp. 61–70 (1999).
59. Andersson, S., Nordstrand, T. and Rasmuson, A., “The Influence of Some Fibre
and Suspension Properties on Pulp Fibre Friction”, J. Pulp Pap. Sci., 26(2): pp.
67–71 (2000).
60. Bennington, C. and Mmbaga, J., “Liquid phase turbulence in pulp fibre suspen-
sions”, in Fundamental Research Symposium, Oxford (2001).
61. Schmid, C. and Kingenberg, D., “Mechanical flocculation in flowing fiber
suspensions”, Phys. Rev. Letters, 84(2): pp. 290–293 (2000).
62. Huhtanen, J.-P., “Non-Newtonian Flows in Paper making”, Energy and Process
Engineering, Tampere University of Technology, Tampere, Techn.lic. Thesis
(1998).
63. Ross, R. and Klingenberg, D., “Simulation of flowing wood fibre suspensions”,
J. Pulp Pap. Sci., 24(12): p. 388 (1998).
64. Schmid, C., Switzer, L. and Klingenberg, D., “Simulation of fiber flocculation:
Effects of fiber properties and interfiber friction”, J. Rheol., 44, pp. 781–809 (2000).
65. Schmid, C. and Klingenberg, D., “Properties of fiber flocs with frictional and
attractive interfiber forces”. J. Coll. Int. Sci., 226, pp. 136–144 (2000).
66. Shah, P., et al., “The role of turbulent elongational stresses on deflocculation in
paper sheet formation”, Tappi J., 83(4): pp. 8 pages (2000).
67. Kaji, H., Monma, K. and Katsura, T., “Fractal analysis of flocculation in pulp
suspension”, in International Paper Physics Conference, Hawaii, pp. 291–297
(1991).
68. Kallmes, O., Kallmes, P. and Bishop, B., “Monitoring flocculation in the paper
machine”, Tappi J., 77(7): pp. 194–198 (1994).
69. Zhao, R. and Kerekes, R., “The effect of suspending liquid viscosity on fiber
flocculation”, Tappi J., 76(2): pp. 183–188 (1993).
70. Paul, T., Duffy, G. and Chen, D., “New insights into the flow of pulp suspen-
sions”, in 54th Appita Annual Conference, Melbourne, pp. 645–654 (2000).
71. Giri, M., Simonsen, J. and Rochefort, W., “Dispersion of pulp slurries using
carboxymethylcellulose”, Tappi J., 83(10): pp. 14 pages (2000).
72. Kerekes, R. and Schell, C., “Effects of fibre length and coarseness on pulp floccu-
lation”, Tappi J., 78(2): pp. 133–139 (1995).
73. Beghello, L., “The tendency of fibers to build flocs”, Laboratory of Paper
Chemistry, Åbo Akademi University, Åbo, PhD Thesis (1998).
74. Raghem-Moayed, A. and Kuhn, D., “Turbulent flocculation measurement”,
J. Pulp Pap. Sci., 26(4): pp. 163–165 (2000).
75. Raghem-Moayed, A., “Characterisation of fibre suspension flows at papermaking
consistencies”, Dept of Chemical Engineering and Applied Chemistry, University
of Toronto, Toronto, PhD Thesis (1999).
76. Wågberg, L. and Nordqvist, T., “Detection of polymer induced flocculation of
cellulosic fibres by image analysis”, Nord. Pulp Pap. Res. J., 14(3): pp. 247–255
(1999).
77. Pierre, C., “A new sensor for the measurement of fibre flocculation in the stock. A
tool for the sheet formation control”, in COST E14, Lisbon (2000).
78. Swerin, A., “Flocculation and fibre network strength in papermaking suspansions
flocculated by retention aid systems”, Division of Paper Technology, KTH,
Stockholm, PhD Thesis (1995).
79. Tuomisaari, M., Gullichsen J. and Hietaniemi, J., “Floc disruption in medium-
consistency fiber suspensions”, in International Paper Physics Conference, Hawai,
pp. 609–612 (1991).
80. Hietaniemi, J. and Gullichsen, J., “Floc disruption in medium consistency fibre
suspensions: Empirical results and their interpretation”, in International Paper
Physics Conference, Niagara-on-the-Lake, pp. P29–38 (1995).
81. Cichoracki, T., Gullichsen, J. and Paulapuro, H., “High consistency forming: A
new concept”, Tappi J., 84(3): pp. 8 pages (2001).
82. Björkman, U., “Flow of flocculated fibres”. Stockholm: Paper Technology, KTH.
283. (1999).
83. Helmer, R., Covey, G. and Lai, L., “Approach flow stock mixing”, in 52nd Appita
Annual Conference, pp. 223–228 (1998).
84. Ringnér, J. and Rasmuson, A., “Characterization of fibre suspensions using X-ray
computed tomography and image analysis”, Nord. Pulp Pap. Res. J., 15(4): pp.
319–325 (2000).
85. Kellomäki, M. et al., “Fiber flocculation measurement in pipe flow by digital
image analysis”, in International Paper Physics Conference, San Diego, pp.
461–463 (1999).
86. Karema, H. et al., “Transient fluidisation of fibre suspension in straight channel
flow”, in International Paper Physics Conference, San Diego, pp. 369–379 (1999).
87. Li, T.-Q., et al., “Velocity measurements of fiber suspensions in pipe flow by the
magnetic resonance imaging method”, Tappi J., 77(3): pp. 145 (1994).
88. Li, T.-Q. and Ödberg, L., “Flow of pulp suspension through an abrupt contrac-
tion studied by flow encoded nuclear magnetic resistance imaging”. Nord. Pulp
Paper Res. J., 10(2): pp. 133 (1995).
89. Li, T.-Q. et al., “Pipe flow behaviour of hardwood pulp suspensions studied by
NMRI”, J. Pulp Pap. Sci., 21(12): pp. J408–414 (1995).
90. Li, T.-Q. and Ödberg, L., “Studies of flocculation in cellulose fibre suspansions by
NMR imaging”, J Pulp Pap. Sci., 23(8): pp. J401 (1997).
91. Houvila, J. et al., “The Valmet Headbox Family-A New Level of Customer
Orientation”, in XII Paper Technology Days, Turku, pp. 128–139 (2000).
92. Pantaleo, S., “Headboxes – Past, Present and Future”, in CPPA Annual Meeting,
Montreal, pp. A131–139 (2000).
93. Schultz, H.-J. “The use of standard units for headbox design”, in 4th
International Conference: New available techniques and current trends, Bologna,
pp. 6–45 (1992).
94. Schultz, H.-J., “Der BTF-Universalverteiler. ein verbindenes Trennelement
zwichen kontasntem Teil und Stoffauflauf”, Wochenbl f. Papierfabr. (11/12): pp.
461–465 (1992).
95. Knoller, H., Foulger, M. and Parisian, J., “A new dilution control system for
existing headboxes”, in 86th Annual Meeting, PAPTAC, Montreal, pp. 17–21
(2000).
96. Begemann, U., “Module Jet, das neue Stoffauflaufkonzept mit Stoffdickte-
Querprofilregelung”, Das Papier, 47(10A): pp. V149 (1993).
97. Begemann, U., “The new headbox concept with CD consistency control”, in
Annual Meeting, Technical Section CPPA, Montreal, pp. A305–310 (1995).
98. Heinzmann, H., “Faserorientierung – Querprofil”, Wochenbl f. Papierfabr., (4):
pp. 121–126 (1995).
99. Pantaleo, S., “A new headbox design featuring consistency profiling decoupled
from fiber orientation response”, Tappi J., 78(11): pp. 89 (1995).
100. Pantaleo, S., “Basis weight profiling experience of the concept IV-MH headbox”,
in Annual Meeting, Technical Section CPPA, Monteal, pp. B193–199 (1996).
101. Nyberg, P. and Malashenko, A., “Dilution control headbox – choises, threats
and solutions”, in 83rd CPPA Annual Meeting, Montreal, pp. 17 (1997).
102. Wurster, H. and Weitkämper, K., “Stoffdichteabhängige Querprofilregelung –
Betriebserfahrungen”, Das Papier, 49(10A): pp. V99–102 (1995).
103. Wyse, R. et al., “Consistency profiling – A new technique for CD basis weight
control”, in Annual Meeting, Technical Section CPPA, Montreal, pp. A267–273
(1995).
104. Heaven, M. et al., “Recent Experiences with Consistency Profile Control Instal-
lations”, in 84th Annual Meeting, Technical Section CPPA, Montreal, pp. 261
(1998).
105. Smits, M. “Results from a consistency profiling headbox and a method of
analysing the CD profile performance”, in 52nd Appita Annual Conference, pp.
397–401 (1998).
106. Pantaleo, S. “A headbox design minimizing MD fiber orientation streaks while
maintaining high resolution basis weight control”, in 85th Annual Meeting,
PAPTAC, Montreal, pp. 113–119 (1999).
107. Collins, D., R. Leclerc and J. Perrault. “Operation of a moduljet dilution
headbox on Dolbeau PM5”, in 85th Annual Meeting, PAPTAC, Montreal, pp.
101–106 (1999).
108. Lee, J. and Pantaleo, S., “Headbox Flow Analysis”, J. Pulp Pap. Sci., 25(12):
pp. 437–441 (1999).
109. Hämäläinen, J. and P. Tarvainen. “CFD-based shape and control optimization
applied to a paper machine headbox”, in 86th Annual Meeting, PAPTAC, Mon-
treal, pp. 99–102 (2000).
110. Fortier, L. “A new family of high turbulence headboxes”, in 76th Annual Meet-
ing, PAPTAC, Montreal, pp. 197–201 (1990).
111. Haptmann, E., Vyse, R. and Mardon, J., “The wake effect as applied to modern
hydraulic headboxes, Parts 1 & 2”, Pulp Paper Can., 91(9: 10): pp. T357–364 and
T369–377 (1990).
112. Shariati, M. et al., “Numerical and experimental models of flow in the con-
verging section of a headbox”, in Tappi Papermakers Conference, Vancouver,
pp. 685–693 (2000).
113. Aidun, C., “Opportunities to improve CD strength and to reduce fibre com-
sumption in boardmaking”, in PIRA International Conference: Scientific and
Technical Advances in Boardmaking, Heathrow (1999).
114. Aidun, C., “Methods and apparatus to enhance paper and board forming
qualities”, IPST patent, US 6 153 057 (2000).
115. Nordström, B., “Effects of headbox design and dewatering conditions on twin-
wire forming of TMP”, Pulp and Paper Chemistry and Technology, KTH,
Stockholm, PhD Thesis (1995).
116. Nordström, B. and Norman, B., “Influence on sheet anisotropy, formation,
Z-toughness and tensile stiffness of reduced feed area to a headbox nozzle”,
Nord. Pulp Pap. Res. J., 9(1): p. 53 (1994).
117. Nordström, B. and Norman, B., “Effects of roll-blade and roll forming on
formation, retention, mecanical properties and anisotropy – A preliminary
comparison”, J. Pulp Pap. Sci., 21(7): pp. J223–229 (1995).
118. Ullmar, M., “On fiber alignment mechanisms in a headbox nozzle”, Pulp and
Paper Chemistry and Technology, KTH, Stockholm, Tekn Lic Thesis (1998).
175. Attwood, B. and Moore, G., “An introduction to the theory and practice of
multiply forming”, Leatherhead: PIRA. 179. (1994).
176. Attwood, B. and Loy, K., “Pressure forming developments in multi-ply paper-
board forming”, Tappi J., 76(10): pp. 160–164 (1993).
177. Halmschlager, G., “Duoformer Top – a new former for the production of top
plies of packaging grades”, in Tappi Engineering Conference, Nashville, pp.
1239–1243 (1997).
178. Räisänen, K. and Paananen, J., “Forming process requirements set by a furnish
type in linerboard production”, in 6th International Conference on New
Available Techniques, Stockholm, pp. 454–457 (1999).
179. Halmschlager, G., “Douformer Base – compact concept for increased board and
packaging paper production”, in 6th International Conference on New Available
Techniques, Stockholm, pp. 460–464 (1999).
180. Norman, B., “Fibre coarseness (weight per unit length) should be replaced by
fibre grammage (weight per unit area)”, Nord. Pulp Paper Res. J., 13(2): p. 166
(1998).
181. Dodson, C., “The effect of fibre length distribution on formation (Research
note)”, J. Pulp Pap. Sci., 18(2): pp. J74–J76 (1992).
182. Dodson, C. and Fekih, K., “The effect of fibre orientation on paper formation”,
J. Pulp Pap. Sci., 17(6): pp. J203–J206 (1991).
183. Dodson, C., “Flocculation and orientation effects on paper formation statistics”,
Tappi J., 75(1): pp. 167–171 (1992).
184. Kuhn, D. and Dodson, C., “Measurement of fibre orientation effects on
formation”, in 78th Annual Meeting, Technical Section CPPA, Montreal,
pp. A283–A286 (1992).
185. Dodson, C. and Serafino, L., “Flocculation, dispersion and dynamic scenarios
for formation”, Nord. Pulp Paper Res. J., 8(2): pp. 264–272 (1993).
186. Deng, M. and Dodson, C., “Random star patterns and paper formation”, Tappi
J., 77(3): pp. 195–199 (1994).
187. Farnood, R., Dodson, C., and Loewen, S., “Modeling flocculation. Part I:
Random disc model”, J. Pulp Pap. Sci., 21(10): pp. J348–355 (1995).
188. Farnood, R. and Dodson, C., “The similarity law of formation”, in Inter-
national Paper Physics Conference, Niagara-on-the-Lake, pp. 5–12 (1995).
189. Farnood, R. et al., “Modelling flocculation: A gallery of simulated flocculated
papers”, Nord. Pulp Paper Res. J., 12(2): pp. 86–89 (1997).
190. Norman, B., “Letter to the editor”, Nord. Pulp Paper Res J., 12(3): p. 210
(1997).
191. Sampson, W. et al., “Hydrodynamic Smoothing in the Sheet Forming Process”,
J. Pulp Pap. Sci., 21(12): pp. J422–426 (1995).
192. Norman, B. et al., “The effect of localised dewatering on paper formation”, in
International Paper Physics Conference, Niagara-on-the-Lake, pp. 55–59 (1995).
193. Lucisano, M. and Norman, B., “The forming and properties of quasi-random
laboratory paper sheets”, in International Paper Physics Conference, San Diego,
pp. 331–340 (1999).
194. Kilpeläinen, R. et al., “Stock Preparation and Wet End. Part 1: forming fabrics”,
in Papermaking Science and Technology, Fapet OY, Helsinki 8(7): pp. 251–283
(2000).
195. Johnson, D., “Effects of jet impingement on Bal Baie machines”, Pulp Paper
Canada, 95(5): pp. 12–15 (1992).
196. Fejér, M., “Der Einfluss des Blattbildungsmediums Sieb auf die Formation”,
Wochenbl. f. Papierfab., (20): pp. 839–846 (1993).
197. Adanur, S., “Effects of forming fabric structural parameters on sheet properties,
Tappi J., 77(10): pp. 187–195 (1994).
198. Hampson, P., “Fibre length and fabric aperture”, in 52nd Appita Annual Confer-
ence, pp. 377–382 (1998).
199. Heinola, M. et al., “The influence of forming fabrics on the retention and drain-
age on gap formers”, in Annual Meeting, Technical Section CPPA, Montreal,
pp. A59–64 (1996).
200. Cole, S., “Huytexx – a new breed of forming fabric”, in 87th Annual Meeting,
PAPTAC, Montreal, pp. A5–7 (2001).
201. Keen, P. and Cole, S., “Unique forming fabric structures deliver added value”, in
87th Annual Meeting PAPTAC, Montreal, pp. A57–60 (2001).
202. Herzig, R. and Johnson, D., “Investigation of thin fiber mats formed at high
velocity”, Tappi J., 82(1): pp. 226–230 (1999).
203. Meyer, H., “The effect of wire screens on forming fiber mats”, Tappi J., 52(9):
pp. 1716–1723 (1969).
204. Danby, R., “The impact of forming fabric structures on print quality”, in CPPA
Annual Meeting, Montreal, pp. 33–38 (1993).
205. Danby, R., “Forming fabric design definition and selection – The means to
machine and fiber efficiency”, in 85th Annual Meeting, PAPTAC, Montreal
(1999).
206. Danby, R. and Plouffe, P., “Print quality improvements through forming fabric
design changes”, Pulp Paper Can., 101(9): pp. 66–69 (2000).
207. Dudley, W. et al., “Engineered approach and unique forming fabric design pro-
vide mill with dramatic paper & print improvements”, in Tappi Engineering
Conference, Atlanta, pp. 9 pages (2000).
208. Jong, J. H., Baines, W. D. and Currie, I. G., “Experimental Characteristics of
Forming Fabrics and Fibre Mats”, J. Pulp Pap. Sci., 25(3): pp. 95–99 (1999).
209. McDonald, D., “Web rewetting by forming fabrics”, in Tappi Engineering
Conference, Anaheim, pp. 597–599 (1999).
210. Barratte, C. et al., “Simulation of initial fibre retention by forming fabric”, in
54th Appita Annual Conference, Melbourne, pp. 415–420 (2000).
211. Keen, P., “Bringing virtual reality to forming fabric design”, Paper Technology
(December): pp. 47–50 (1999).
212. Sayegh, N. and Gonzales, T., “Compressibility of Fibre Mats During Drainage”,
J. Pulp. Pap. Sci., 21(7): pp. J255–261 (1995).
213. Vomhoff, H. and Schmidt, A., “The steady-state compressibility of saturated
finre webs at low pressure”, Nord. Pulp Pap. Res. J., 12(4): pp. 267–269 (1997).
of fine paper”, in 4th New Available Techniques and Current Trends Conference,
Bologna, pp. 46–72 (1992).
233. Kiviranta, A. and Paulapuro, H., “Characterization and optimization of
Fourdrinier table activity in linerboard manufacture”, Paperi ja Puu, 74(9): pp.
728–737 (1992).
234. Kiviranta, A. and Paulapuro, H., “The role of fourdrinier table activity in the
manufacture of various paper and board grades”, Tappi J., 75(4): pp. 172–186
(1992).
235. Ahonen, P., Kiviranta, A. and Kaipila, J., “Effect of Fourdrinier table layout on
the wet-end performance of a fine paper machine”, Paperi ja Puu, 74(10): pp.
802–806 (1992).
236. Foulger, M., “Submerged drainage: A new concept of dewatering”, in Annual
Meeting, Technical Section CPPA, Montreal, pp. A65–A68 (1995).
237. Zu, J. and Chen, J., “A theoretical study of vacuum force on gravity foils and
step foils”, Tappi J., 82(11): pp. 93–98 (1999).
238. Evans, D. and Fielding, S., “VID formation system”, in 54th Appita Annual
Conference, Melbourne, pp. 437–442 (2000).
239. Palm, D. and Sodergren, O., “VID dormation system”, in 87th Annual Meeting
PAPTAC, Montreal, pp. A185–190 (2001).
240. Sodergren, O. and Neun, J., “Developments in activity generation on
Fourdriniers”, Tappi J., 83(10): pp. 10 pages (2000).
241. Verkasalo, L., Odell, M. and Korhonen, H., “Forming of LWC base papers”, in
Tappi Engineering Conference, Nashville, pp. 1231–1238 (1997).
242. Bando, T., Iwata, H. and Nagano, A., “Drainage mechanism on a twin-wire
former. Part 1: Factors affecting on the drainage phenomena”, Japan Tappi,
48(7): pp. 948–954 (1994).
243. Bando, T., Adachi, T. and Iwata, H., “Drainage mechanism on a twin-wire
former. Part 2: Computer simulation on pulsating drainage”, Japan Tappi,
48(11): pp. 1493–1498 (1994).
244. Iwata, H. et al., “Development of Mitsubishi new former for papermaking
machine”. Mitsubishi Heavy Industries, Technical Review, Technical Review,
Vol. 30, No. 2, 123–128 (1993).
245. Zhao, R. and Kerekes, R., “Pressure distribution between forming fabrics in
blade gap formers: Thin blades”, J. Pulp Pap. Sci., 21(3): pp. 97–103 (1995).
246. Zhao, R. and Kerekes, R., “The effect of consistency on pressure pulses in blade
gap formers”, Paperi Ja Puu, 78(1/2): pp. 36–38 (1996).
247. Nigam, M. and Bark, F., “An analytical method to calculate the flow past a
blade in twin-wire formers”, Dep of Mechanics, KTH, Stockholm, 1997: 7
(1997).
248. Zahrai, S., “On the fluid mechanics of twin-wire formers”, Dep of Mechanics,
KTH, Stockholm, PhD Thesis (1997).
249. Zahrai, S. and Bark, F., “On the fluid mechanics of twin wire blade forming in
paper machines”, Nord. Pulp Pap. Res. J., 10(4): pp. 245–252 (1995).
250. Zahrai, S. and Bark, F., “Analytical and numerical approaches to solve the
model equations for the flow of pulp in blade forming process”, in AIChE
Symposium Series, Pulping, Papermaking and Chemical Preparation, pp. 68–73
(1996).
251. Zahrai, S., Bark, F. and Norman, B., “An analysis of blade dewatering in a
twin-wire paper machine”, J. Pulp Pap. Sci., 23(9): pp. J452–459 (1997).
252. Green, S., “Analytical and computational modeling of twin-wire blade forming”,
J. Pulp Pap. Sci., 23(7): pp. J353–358 (1997).
253. Roshanzamir, A., Green, S. and Kerekes, R., “Two-Dimensional Simulation of
Pressure Pulses in Blade Gap Formers”, J. Pulp Pap. Sci., 24(11): pp. 364–368
(1998).
254. Green, S., Zhao, R. and Kerekes, R., “Pressure Distribution Between Forming
Fabrics in Blade Gap Formers: Blades of Finite Width and Fabrics of Finite
Stiffness”, J. Pulp Pap. Sci., 24(2): pp. 60–66 (1998).
255. Green, S. and Kerekes, R., “Numerical analysis of pressure pulses induced by
blades in gap formers”, Tappi J., 81(4): pp. 180–187 (1998).
256. Roshanzamir, A. et al., “Hydrodynamic Pressure Generated by Doctoring in
Bade Gap Formers”, in Tappi Engineering Conference, pp. 1181–1189 (1999).
257. Roshanzamir, A., Green, S. and Kerekes, R., “The Effect of Non-Darcy’s Law
Drainage on the Hydrodynamics of Blade Gap Formers”, in Tappi Papermakers
Conference, Vancouver, pp. 701–708 (2000).
258. Shands, J. and Wildfong, V., “A twin wire former rebuild option for improved
formation and drainage”, in Tappi Engineering Conference, pp. 53–69 (1998).
259. Odell, M., Pakarinen, P. and Luontama, M., “Customizing roll and blade form-
ing to control paper structure”, in Valmet Paper Machine Days, Jyväskylä, pp.
21–42 (1996).
260. Green, S., “Modeling suction shoes in twin-wire blade forming: Theory”, Tappi
J., 82(9): pp. 136–142 (1999).
261. Green, S., “Modelling Suction Shoes in Twin-Wire Forming: Results”, J. Pulp
Pap. Sci., 26(2): pp. 53–58 (2000).
262. Roshanzamir, A., Green, S. and Kerekes, R., “Two-Dimensional Simulation of
Suction Shoes in Gap Formers”, J. Pulp Pap. Sci., 26(4): pp. 158–162 (2000).
263. Martinez, M., “Characterizing the dewatering rate in roll gap formers”, J. Pulp
Pap. Sci., 24(1): pp. 7–13 (1998).
264. Boxer, T., “Quantitative modelling of an industrial roll-blade gap former”,
Department of Paper Science, UMIST, Manchester, PhD Thesis (1999).
265. Boxer, T., Dodson, C. and Sampson, W., “Analytic solution to the Martinez
dewatering equations for roll gap formers”, J Pulp Pap. Sci., 26(11): pp. 391–394
(2000).
266. Zahrai, S., Martinez, M. and Dahlkild, A., “Estimating the thickness of the web
during twin-wire forming”, J. Pulp Pap. Sci., 24(2): pp. 67–72 (1998).
267. Zahrai, S., Bark, F. and Martinez, M., “A numerical study of cake formation in
2-D cross-flow filtration”, J. Pulp Pap. Sci., 24(9): pp. 281–285 (1998).
268. Turnbull, P. et al., “One-dimensional dynamic model of a paper forming
process”, Tappi J., 80(1): pp. 245–252 (1997).
269. Chen, E., Schultz, W. and Perkins, N., “A two-dimensional viscous model
of a wet paper forming process”, in Tappi Engineering Conference, pp. 21
(1998).
270. Malashenko, A. and Karlsson, M., “Twin Wire Forming – An Overview”, in
Annual Meeting PAPTAC, Montreal, pp. A189–201 (2000).
271. Wildfong, V. et al., “Drainage during roll forming – Model validation using pilot
papermachine data”, in Tappi Engineering Conference, Atlanta, pp. 11 pages
(2000).
272. Gooding, R., McDonald, D. and Rompré A., “Measurement of drainage
around a forming roll”, in 87th Annual Meeting, PAPTAC, Montreal, pp.
A125–138 (2001).
273. Neun, J., “Performance of high vacuum dewatering elements in the forming
section”, Tappi J., 77(9): pp. 133–138 (1994).
274. Neun, J. and Fielding, S. “High vacuum dewatering optimization”, in Tappi
Papermakers Conference, pp. 307–312 (1994).
275. Neun, J., “High vacuum dewatering of brown paper grades”, in Tappi Paper-
makers Conference, pp. 259–265 (1995).
276. Neun, J., “High-vacuum dewatering of newsprint”, Tappi J., 79(9): pp. 153–157,
(1996).
277. Gagnon, J. and Neun, J., “High vacuum dewatering on fourdriniers and for-
mers”, in Annual Meeting, Technical Section CPPA, Montreal (1996).
278. Räisenen, K., “Water removal by flat boxes and a couch roll on a paper machine
wire section”, Laboratory of Paper Technology, HUT, Helsinki, PhD Thesis
(1998).
279. Räisänen, K., “High-vacuum dewatering on a paper machine wire section – a
literature review”, Paperi ja Puu (Papper och Trä), 78(3): pp. 113–120 (1996).
280. Räisänen, K., Paulapuro, H. and Karrila, S., “The effects of retention aids,
drainage conditions and pretreatment of slurry on high-vacuum dewatering; a
laboratory study”, Tappi J., 78(4): pp. 140–147 (1995).
281. Räisänen, K., Karrila, S. and Maijala, A., “Effect of vacuum level and suction
time on vacuum assisted drainage of a paper machine wire section”, Appita J.,
48(4), pp. 269–274 (1995).
282. Räisänen, K., Karilla, S. and Maijala, A., “Vacuum dewatering optimization
with different furnishes”, Paperi ja Puu (Papper och Trä), 78(8): pp. 461–467
(1996).
283. Mitchell, C. and Johnston, R., “Pulsating suction during vacuum dewatering
and its effect on the rate and extent of water removal”, in 54th Appita Annual
Conference, Melbourne, pp. 443–447 (2000).
284. Shands, J. and Hardwick, C., “Dewatering on a high vacuum flat box”, in 86th
Annual Meeting, PAPTAC, Montreal, pp. A113–117 (2000).
285. Jones, G., “Dynamic simulation of dewatering in high vacuum flat boxes”, in
Tappi Engineering Conference, Anaheim, pp. 197–209 (1999).
286. Nazhad, M. and Kerekes, R., “Producing handsheets at headbox consistency”,
in Annual Meeting, PAPTAC, Montreal, pp. A203–208 (1999).
307. Albinsson, C.-J., Swerin, A. and Ödberg, L., “Formation and retention during
twin-wire blade forming of a fine paper stock”, Tappi J., 78(4): pp. 121–128
(1995).
308. Swerin, A. and Mähler, A., “Formation, retention and drainage of a fine paper
stock during twin-wire roll-blade forming. Implications of fibre network
strength”, Nord. Pulp Paper Res. J., 11(1): pp. 36–42 (1996).
309. Nilsson, B., “Fibre orientation effects in twin wire forming”, (in Swedish). Dep.
Paper Technology, KTH, Stockholm, Licentiate Thesis (1993).
310. Eguchi, A. et al., “Flow characteristics in a headbox and fiber orientation of
paper. Part 2: Verification by the experimental paper machine”, Japan Tappi,
47(7): pp. 887–889 (1993).
311. Nordström, B. and Norman, B., “Effects on formation, retention and mechan-
ical properties of forming blade distance in a twin-wire roll-blade former”,
Appita J., 48(1): pp. 19–24 (1995).
312. Nordström, B. and Norman, B., “Effects of headbox jet quality and blade pulse
force on formation, retention and mechanical properties during roll-blade form-
ing with curved blades”, Nord. Pulp Pap. Res. J., 10(1): pp. 33–40 (1995).
313. Nordström, B. and Norman, B., “Effect on paper properties and retention of the
proportion of roll dewatering during twin-wire roll-blade forming of TMP”,
J. Pulp Pap. Sci., 22(8): p. J283 (1996).
314. Odell, M., “Paper structure engineering”, in 53rd Appita Annual Conference,
Rotorua, pp. 155–160 (1999).
315. Erkkilä, A.-L., Pakarinen, P. and Odell, M., “The effect of forming mechanisms
on layered fiber structure in roll and blade gap forming”, in Tappi Papermakers
Conference, pp. 389–400 (1999).
316. Mohlin, U.-B., “Pulp evaluation for modern papermaking”, in 6th International
Conference on New Available Techniques, Stockholm, pp. 343–350 (1999).
317. Mohlin, U.-B., “Fibre dimensions-formation and strength”, in International
Paper Physics Seminar, Grenoble, pp. 5–14 (2000).
318. Odell, M. and Pakarinen, P., “The complete fibre orientation control and effects
on diverse paper properties”, in Tappi Papermakers Conference, Cincinatti, pp.
27 pages (2001).
319. Page, R. and Hergert, R., “Enhancement of Paper Properties and Fibre
Economics Through Web Stratification”, Appita J., 42(1): pp. 33–41 (1989).
320. Häggblom-Ahnger, U., “Three-ply office paper”, Laboratory of Paper Chemistry,
Åbo Akedemi University, Åbo, PhD Thesis (1998).
321. Häggblom-Ahnger, U. and Eklund, D., “The role of the middle ply in multilay-
ered copy paper”, Appita Journal, 51(3): pp. 199–204 (1998).
322. Häggblom-Ahnger, U., “Optimum location of softwood sulphate pulp in
three-ply office paper”, Tappi J., 82(6): pp. 181–187 (1999).
323. Häggblom-Ahnger, U., “Layering of office paper”, Paper and Tiber, 80(7): pp.
508–513 (1998).
324. Lloyd, M., “Stratified (multilayer) forming – a technology for the new millenium?
in 53rd Appita Annual Conference, Rotorua, pp. 161–168 (1999).
Bo Norman
Well our basic finding is that if you do these improvements in the head box
such that you have an improved jet coming out then you will get better
strength. If you introduce some of these shear influences in dewatering you
will lose strength. So when stock is coming onto the wire section it is too late
to make any strength improvements; these have to be upstream in the head
box. We think that is one conclusion.
Bo Norman
I think that the hybrid former is a retro-fit for older machines. I do not think
that you would fit this technology to new machines. For an optimum machine
you should not combine Fourdrinier dewatering forming and twin wire
dewatering, that will not be the best option.
Jean-Claude Roux
Why do we do it like that?
Bo Norman
We want to dewater paper under minimum shear. If you start dewatering on a
Fourdrinier wire and you add a top wire the suspension will enter into a
pressure zone that will decelerate the suspension which will inevitably cause
shear. This should be avoided in the final, perfect paper machine but we have
a long way to go so with today’s machines, of course for rebuilds on a
Fourdrinier machine you will often have a positive formation effect when
introducing these top wires.
Jean-Claude Roux
You mean we lose orientation profile at the beginning of the hybrid former,
and it is not possible to achieve it later in the process?
Bo Norman
Yes, the best alternative is however to add stationary and adjustable blades in
the top wire unit, to be able to control and minimize shear effects.
Session 3