21
This is an electronic reprint of the original article. This reprint may differ from the original in pagination and typographic detail. Author(s): Hauru, Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert Title: Dry jet-wet spinning of strong cellulose filaments from ionic liquid solution Year: 2014 Version: Author accepted / Post print version Please cite the original version: Hauru, Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert. 2014. Dry jet-wet spinning of strong cellulose filaments from ionic liquid solution. Springer Netherlands, Cellulose, volume 20, issue 6, pages 4471-4481. ISSN 1572-882X. DOI: 10.1007/s10570-014-0414-0 Rights: © 2014 Springer. This is author accepted / post print version of the article published by Springer "Hauru, Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert. 2014. Dry jet-wet spinning of strong cellulose filaments from ionic liquid solution. Springer Netherlands, Cellulose, volume 20, issue 6, pages 4471-4481". This publication is included in the electronic version of the article dissertation: Hauru, Lauri K.J. Lignocellulose solutions in ionic liquids. Aalto University publication series DOCTORAL DISSERTATIONS, 87/2017. All material supplied via Aaltodoc is protected by copyright and other intellectual property rights, and duplication or sale of all or part of any of the repository collections is not permitted, except that material may be duplicated by you for your research use or educational purposes in electronic or print form. You must obtain permission for any other use. Electronic or print copies may not be offered, whether for sale or otherwise to anyone who is not an authorised user. Powered by TCPDF (www.tcpdf.org)

Dry jet-wet spinning of strong cellulose filaments from

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Dry jet-wet spinning of strong cellulose filaments from

This is an electronic reprint of the original article.This reprint may differ from the original in pagination and typographic detail.

Author(s): Hauru, Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert

Title: Dry jet-wet spinning of strong cellulose filaments from ionic liquidsolution

Year: 2014

Version: Author accepted / Post print version

Please cite the original version:Hauru, Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert. 2014. Dry jet-wet spinningof strong cellulose filaments from ionic liquid solution. Springer Netherlands, Cellulose,volume 20, issue 6, pages 4471-4481. ISSN 1572-882X. DOI:10.1007/s10570-014-0414-0

Rights: © 2014 Springer. This is author accepted / post print version of the article published by Springer "Hauru,Lauri; Hummel, Michael; Michud, Anne; Sixta, Herbert. 2014. Dry jet-wet spinning of strong cellulosefilaments from ionic liquid solution. Springer Netherlands, Cellulose, volume 20, issue 6, pages 4471-4481".

This publication is included in the electronic version of the article dissertation:Hauru, Lauri K.J. Lignocellulose solutions in ionic liquids.Aalto University publication series DOCTORAL DISSERTATIONS, 87/2017.

All material supplied via Aaltodoc is protected by copyright and other intellectual property rights, andduplication or sale of all or part of any of the repository collections is not permitted, except that material maybe duplicated by you for your research use or educational purposes in electronic or print form. You mustobtain permission for any other use. Electronic or print copies may not be offered, whether for sale orotherwise to anyone who is not an authorised user.

Powered by TCPDF (www.tcpdf.org)

Page 2: Dry jet-wet spinning of strong cellulose filaments from

1

Dry jet-wet spinning of strong cellulosefilaments from ionic liquid solution

Lauri K. J. Hauru, Michael Hummel, Anne Michud, Herbert Sixta

Department of Forest Products Technology, Aalto University, School of ChemicalEngineering, P.O. Box 16300, 00076 Aalto, Espoo, FinlandPhone: +358-50-384 1764

Fax: +358-9-470 24259

E-mail: [email protected]

Considerable growth is expected in the production of man-made cellulose textile fibers, which are

commercially produced either via derivatization to form cellulose xanthate (viscose) or via direct

dissolution in N-methylmorpholine N-oxide (Lyocell). In the study at hand, cellulosic fibers are

spun from a solution in the ionic liquid [DBNH][OAc] into water, resulting in properties equal or

better than Lyocell (tensile strength 37 cN tex–1 or 550 MPa). Spinning stability is explored, and

the effects of extrusion velocity, draw ratio, spinneret aspect ratio and bath temperature on

mechanical properties and orientation are discussed. With the given set-up, tenacities and moduli

are improved with higher draw ratios, while elongation at break, the ratio of wet to dry strength,

modulus of resilience and birefringence depend little on draw ratio or extrusion velocity, elastic

limit not at all. We find the process robust and simple, with stretching to a draw ratio of 5 effecting

most improvement, explained by the orientation of amorphous domains along the fiber axis.

textile fibers, spinning, extrusion, regeneration, tenacity, modulus

Introduction

The global demand for textile fibers is expected to increase following population

growth and improvement in standard of living. Due to the inherent properties of

cellulosic fibers they will cover 33–37% of the total fiber market, resulting in an

increase of per-capita demand from the current 3.7 kg to 5.4 kg in 2030. Cotton

production cannot be increased significantly, creating a 1.7 kg per capita

"cellulose gap" for man-made cellulosic fibers. With the present capacities, there

will be an additional demand of 15 million t of cellulosic fibers in 2030

(Hämmerle 2011). Currently, most man-made cellulosic fibers are produced by

the viscose process, with minor shares from Lyocell (3.8% or 222 000 tons out of

a 5.8 million ton market) and cupro (special applications) (Happel 2014). In the

viscose process, cellulose is xanthogenated, dissolved in caustic soda, and wet

Page 3: Dry jet-wet spinning of strong cellulose filaments from

2

spun in an aqueous bath containing sulfuric acid, sodium sulfate and additives that

delay the regeneration. In the Lyocell process, cellulose is dissolved in a direct

solvent (N-methylmorpholine N-oxide, NMMO) to give a solution (termed dope)

and dry jet-wet spun through an air gap into an aqueous coagulation bath. The

mechanical properties of Lyocell filaments are significantly better than those of

regular viscose filaments: e.g. the tensile strengths are 630 MPa vs. 330 MPa,

respectively (Fink et al. 2001; Röder et al. 2009). The production of NMMO-

based Lyocell fibers has been thoroughly investigated with respect to the effects

of spinning variables (Mortimer and Péguy 1996a; Mortimer et al. 1996;

Mortimer and Péguy 1996b).

In filament spinning, a molten polymer or a solution of a polymer is extruded

through a die (spinneret) with a certain extrusion velocity (expressed either as a

velocity in m/min or as a volumetric flow rate in ml/min) and drawn with rotating

godets. The take-up godet is set to a higher linear velocity than the extrusion

velocity to develop the targeted tenacity by stretching the filament. The ratio of

velocities is termed the draw ratio (DR). Polymer molecules are rotated towards

the axial direction, pulled, untangled and straightened, thus producing a highly

oriented structure. In cellulose filaments, orientation promotes crystallization and

produces a dense, continuous microfibrillar structure that has a high tensile

strength. Crystallinity and crystalline orientation is measured with wide-angle X-

ray scattering, while the total orientation is observed as an increase in

birefringence. The difference of total and crystalline orientation is denoted as the

amorphous orientation, which is strongly related to the tensile strength. In

polymers, relaxation is slow, thus structure, including axial orientation of even

amorphous material, is retained after mechanical deformation of a solution or

melt, and may be set by solidification. Thus, the extruded and drawn jet is quickly

solidified by cooling (melt spinning), evaporation of the solvent (dry spinning), or

by precipitation with an antisolvent (wet spinning) (Stibal et al. 2000). Together,

crystallite size (measured with level-off degree of polymerization), molar mass,

total orientation, and crystallinity determine the strength properties in excess of

those in an unoriented, undrawn state. (Krässig 1993)

Page 4: Dry jet-wet spinning of strong cellulose filaments from

3

In the viscose process, dopes with relatively low viscosity (zero shear viscosity

about 5–20 Pas) are used, and the dope is extruded directly to the bath (wet

spinning) to immediately form a regenerated "skin" layer on the surface, after

which sulfuric acid diffuses in and solidifies the interior. In Lyocell, by contrast,

highly viscous dopes (zero shear viscosity in the range of 5 000–30 000 Pas at

spinning temperature) are produced. The dope is extruded into an ca. 1–5 cm air

gap, where drawing and cooling is accomplished, and then the filament

immediately enters a coagulation bath, where the structure is set uniformly by a

spinodal decomposition-type phase separation. Neutron diffraction studies have

shown that the fiber cross section consists of micro- and nanofibrils of uniform

dimensions surrounded by a very thin fiber skin. (Schuster et al. 2003) In this dry

jet-wet spinning process, high orientation is accomplished by stretching the

solution before regenerating it. Whereas, in the regular viscose process, relatively

amorphous, low-tenacity fibers are spun with draw ratios less than 1, and

stretching is done after the regeneration (Hearle and Schawaller 2000; Stibal et al.

2000; Fink et al. 2001; Coulsey and Smith 1996). Hence, potentially, higher

viscosities and thus lower temperatures would be preferable. However, these are

limited by the corresponding increase of viscosity of the solution and higher

extrusion pressure, and the melting points of solvents. In the coagulation bath,

higher temperatures would promote diffusion and rapid solidification, but would

allow the polymers to relax and reduce the strength of the forming filament.

There has been considerable research interest in ionic liquids (ILs) during the past

decade, with numerous applications proposed. ILs with high hydrogen bonding

basicities dissolve cellulose, which can be then regenerated by reducing the net

hydrogen bonding basicity (β–α) of the solution with water (Swatloski et al. 2002;

Doherty et al. 2010; Hauru et al. 2012a). Thus, dopes can be prepared and spun

from suitable ionic liquids such as the 1-butyl-, 1-allyl or 1-ethyl-3-

methylimidazoliums [bmim]Cl, [amim]Cl or [emim]OAc (Viswanathan et al.

2006; Sun et al. 2008; Zhang et al. 2007). We have reviewed these systems in

detail (Hummel et al. 2014). ILs have several advantages; the chemical inertness

of ILs if compared to the viscose reagents (NaOH/CS2) or the Lyocell solvent (a

N-oxide) reduces chemical hazards and allows functional additives that would

react with other solvents. However, these are not without their own problems.

Page 5: Dry jet-wet spinning of strong cellulose filaments from

4

Chloride-containing ILs such as [bmim]Cl cause corrosion problems in industrial

equipment, and can be environmentally toxic (Bernot et al. 2005). Imidazolium

ILs are quaternary ammonium compounds that, once formed, remain liquids up to

very low pressures and are exceedingly difficult to purify, as distillation requires

high vacuum (0.03–0.05 mbar) and a high temperature (175 ºC) and causes

dealkylation and other degradation of the IL (King et al. 2012). Imidazoliums are

also not chemically inert: deprotonation at the C-2 position gives a reactive N-

heterocyclic carbene. It can attack the reducing end of the polysaccharide, and as a

nucleophilic catalyst, causes depolymerization (Grasa et al. 2002; Schrems et al.

2010). Also, depolymerization occurs, as the proportion of the longest chains

(DP>2000) is significantly reduced when cellulose is dissolved in [emim]OAc.

Most importantly, however, the tensile strengths achieved in such processes are

typically 130–375 MPa (9–25 cN/tex), which is inferior to Lyocell, but similar to

viscose. In other examples with sufficiently high cellulose concentration and draw

ratio, reasonable (402–569 MPa or 27–38 cN/tex, Laus et al. (2005)) or good

tenacities (654–801 MPa or 44–53 cN/tex, Kosan et al. (2008)) were obtained

with alkylmethylimidazolium chlorides or [emim]OAc. We have also studied the

use of [emim]OAc for spinning. However, we have only attained a tenacity of 13–

16 cN/tex (200–240 MPa) with a maximum DR of 3–4 for 6–20% solutions, and

only so if spun from a large 250 µm spinneret (Hummel et al. 2011; Hauru et al.

2012b).

In this study, we present spinning from a novel ionic liquid solvent, 1,5-

diazabicyclo[4.3.0]non-5-enium acetate ([DBNH][OAc]) (Michud et al. 2014;

Hummel et al. 2014; Parviainen et al. 2013). This solvent solves many of the

aforementioned problems: it is distillable, forms dopes with a rheology practically

identical to that of NMMO, requires only a cold water bath for cellulose

regeneration, is noncorrosive and safe to handle, and most importantly, gives

exceptionally strong filaments, reaching or exceeding the level achieved in

Lyocell. Recycling of [DBNH][OAc] in the spinning process is expected to be

very similar to that in the Lyocell process: after treatment to remove impurities,

the solvent is simply reconcentrated by evaporation of water. The effects of draw

ratio, extrusion velocity, spin bath temperature and spinneret aspect ratio on

mechanical properties and orientation are explored by experiment.

Page 6: Dry jet-wet spinning of strong cellulose filaments from

5

Materials and methods

[DBNH][OAc] was synthesized by protonation of DBN (1,5-

diazabicyclo[4.3.0]non-5-ene, 99%, Fluorochem, UK) with acetic acid (99.8%).

The acid was added carefully in increments, allowing the temperature to rise to 80

ºC to keep the forming IL in liquid state. The stoichiometry was checked with 1H

NMR analysis with a Varian Unity Inova 600 spectrometer (King et al. 2011).

Prehydrolysis-kraft dissolving pulp (Eucalyptus urograndis, 93% cellulose I, Mw

= 269 kg mol–1, Mn = 79 kg mol–1 and polydispersity 3.4, Bahia Speciality

Cellulose, Brazil) was cut into powder in a Wiley mill (1 mm sieve). The dope

(13% cellulose) was prepared with a vertical kneader system with an anchor blade

as described in detail previously (Hauru et al. 2013). Air-dry pulp was added to

the molten IL in 2 m-% increments, while kneading for 3 h at 10 rpm and 80 ºC in

50–200 mbar vacuum to remove air bubbles due to entrained air from the pulp.

The solution was press-filtered with 2 MPa at 100 ºC through layered filter mesh

(GKD Ymax2, 5 µm nominal, Gebr. Kufferath AG, Germany), deaerated

overnight at 80 ºC in a vacuum oven (<50 mbar), and stored protected from

moisture in double heat-sealed polyethylene bags.

To determine the rheological properties of the dope, a frequency sweep was done

at spinning temperature (70 °C) with a Paar Physica MCR 300 rheometer, with a

plate-plate geometry (25 mm plate, 1 mm measuring gap). The zero-shear

viscosity from the Cross model (Cross 1965) was 11551 Pa·s. The cross-over

point was at the angular frequency of 2.31 s–1, where the moduli were 5315 Pa.

The dope exhibits the expected shear thinning behavior (Fig S1), and is in terms

of rheology practically identical to a corresponding 13% NMMO dope at 95 °C in

the high-frequency regime.

Filament spinning was done with a laboratory piston spinning system (Fourné

Polymertechnik, Germany) through monofilament spinnerets with a diameter of

100 µm and a length-diameter ratio of the spinneret capillary (L/D) of 0.2 or 2

(Enka Tecnica, Germany). Cylinder pressure was measured with a sensor and

limited with a torque-limiting coupling. The spinning cylinder was heated to 70 ºC

and loaded with a piece of the solidified dope (10 g), softened by slight heating,

Page 7: Dry jet-wet spinning of strong cellulose filaments from

6

by means of a syringe-type applicator. The filaments were spun over a 1 cm air

gap into a cold water bath (15 ºC), where the formed filament was led over a

teflon guide roller (at 20 cm depth) and via another guide onto a godet couple.

The extrusion velocity (ve) was varied between 0.01 and 0.045 ml min–1 (1.27–

5.73 m/min), and the draw ratio (DR) varied between 1 and 12.5 by adjusting the

take-up speed (5–38 m min–1). The minimum godet speed was 5 m min–1, thus DR

1 was outside the envelope for 0.02 ml min–1 and only DR 1.3 was attainable at

0.03 ml min–1. The filaments were collected on the godet, carefully cut with a

razor blade and removed, washed first in cold (5 ºC) and then in hot water (50 ºC,

5 min), and dried and conditioned without tension in a controlled atmosphere (23

ºC, 50% RH).

Titer determination and tensile testing in dry and wet mode were carried out with

a Vibroskop–Vibrodyn system (Lenzing Instruments GmbH & Co KG, Austria) in

23 ºC and 50% RH. 10 samples were tested, the gap length was 20 mm and speed

10 mm s–1, and a pretension of 5.9±1.2 mN tex–1 was used (BISFA 2004). Breaks

at <5% (dry) and <10% (wet) elongation were considered incidental and ignored.

The proportional limits and initial moduli were calculated from the stress-strain

curves with a MatLab script according to ASTM standard D2256/D2256M; see

supporting information for details (ASTM 2010). All specimens had clearly

distinguishable elastic and plastic regions in their stress-strain curves when dry,

and only a single region when wet. Resilience (elastic energy density before yield)

was calculated as:

(1)

as yield stress and U is defined as , where U is resilience

(MJ m–3), E is Young's modulus (in MPa), εp is strain at proportional limit

(dimensionless) and σp is stress at proportional limit (MPa).

Birefringence was determined from three selected filaments with a Zeiss Axio

Scope A1 microscope and a Leica B 5λ-Berek tilting compensator. Titer was

measured three times with the vibroscope, permitting 1.5% variation to exclude

Page 8: Dry jet-wet spinning of strong cellulose filaments from

7

artifacts, and the diameter calculated assuming a density of 1.5 g ml–1. The

filament was tensioned between two pieces of double-sided tape on a microscope

slide. The optical retardation was determined in triplicate from a selected spot.

Birefringence is the retardation divided by the diameter; a birefringence of 0.062

was assumed to be equivalent to 100% orientation (Lenz et al. 1994).

Results and discussion

A spinning envelope (Fig. 1) for this system was constructed by finding the

extrusion velocities and draw ratios where spinning is possible, in DR steps 1, 3, 5,

7.5, 10 and 12.5, using a spinneret with an length-diameter ratio of the capillary

(L/D) of 2. For Fig. 1, spinning was defined as stable if continuous take-up was

possible for 1–2 minutes without breakage. (This was a practical choice; proper

stability testing for industrial use would involve a multifilament spinneret.) Due to

limitations in take-up speed (minimum godet speed: 5 m min–1) not all draw

ratios were accessible at low extrusion velocities, which is indicated by the white

area in Fig. 1. Only DR 2 was possible at 0.02 ml min–1 and DR 1.3 at 0.03 ml min–

1. The extrusion velocities were limited to 0.045 ml min–1. At higher throughput

(purple area in Fig. 1) fluctuations of the cylinder pressure result in critical

overpressure (33–41 bar). At 0.01 ml min–1, take-up was difficult and spinning not

very stable, with breaks in the bath (yellow area). Filament breaks between the

second guide and the godet limited draw ratios in an extrusion velocity-dependent

manner, with limits of DR 7.5–12.5 (30–40 m min–1). According to the usual

classification into cohesive, capillary and telescope breaches (Gries et al. 2011),

this is a cohesive fracture. At the breakpoint, the filament has solidified, thus it

cannot transmit capillary waves, and does not have the liquid core a telescope

breach requires.

Page 9: Dry jet-wet spinning of strong cellulose filaments from

8

Fig. 1 The spinning envelope, with circles for sampled points; white dashed lines indicate

takeup velocities

Linear density (titer)

In order to determine the function describing the dependency of the titer from the

draw ratio, the titer (linear density, d) was measured as a function of the draw

ratio for L/D 2 at extrusion velocity ve = 0.02 ml min–1 and L/D 0.2 at 0.04 ml

min–1. There was no major difference between the two, and combined, they are

well described (Fig. 2) by a reciprocal function: d = (22.4±0.4 dtex) DR–1.

Permitting the negative exponent to vary, it was 0.95±0.02, i.e. consistent with 1,

which is in contrast to 0.5 found by Kong and Eichhorn and Mortimer for NMMO

(Kong and Eichhorn 2005; Mortimer and Péguy 1996a). For NMMO, the change

of extensional viscosity with cooling of the jet results in an exponent other than

unity. Instead, here, a mass transport balance for simple deformation describes the

draw:

(2)

where ρ is dope density (1100 kg m–³), w polymer concentration (13 m-%), r

spinneret radius (50 µm) and s a correction factor for contraction on solidification

and drying. ρwπr² = 11.2 dtex, thus s = 1.994±0.004. This factor has the same

value as in the Lyocell process. (See supporting information for further

discussion.)

Page 10: Dry jet-wet spinning of strong cellulose filaments from

9

Fig. 2 Titer as a function of draw ratio, with aspect ratios and guide rotation varied; gray lines:

confidence and prediction bands

Titers from varying extrusion velocity (Ve) fit to the same form of equation: d =

(23.1±0.2 dtex) DR–1 (Fig. 3), showing that titer was unaffected by extrusion

velocity and thus absolute speed. In contrast, Mortimer et al. (Figure 2, p. 254–

255) found a large effect from extrusion velocity on jet profile with NMMO. In

that system (15% solution extruded at 115 ºC, 150–200 µm spinneret, 0.041–

0.096 ml/min, DR 3.6–15.4), the filament profile was set by solidification due to

cooling. Due to this, a higher extrusion velocity produced significantly thicker jets

up to an air gap of 100 mm. Line speed itself changed the jet profile only a little

due to slightly more efficient cooling at lower speeds (Mortimer et al. 1996).

Another contrasting system was 5% cellulose in [bmim]Cl at 90 ºC (spinneret 150

µm, 0.14–0.22 ml/min or 7.7–12.3 m/min), for which Jiang et al. (2012) found a

significant change of properties with speed. However, they used a relatively

nonviscous dope (zero-shear viscosity 50 Pas) for which other forces than

resistance to deformation, such as the inertial contribution or skin drag, could be

important (see supporting information). Whereas, in our system an extruded

volume element undergoes only simple deformation in the jet.

Page 11: Dry jet-wet spinning of strong cellulose filaments from

10

Fig. 3 Titer vs. draw ratio and extrusion velocity (ml min–1)

Tenacity and modulus

As expected, tenacity (specific tensile strength, σ) increases with higher draw

(Fig. 4), reaching 36.8±2.2 cN tex–1 (552±34 MPa) at DR ≥ 5, which is similar to

Lyocell. At 18.9–19.9 cN tex–1 (310±33 MPa), the tenacity of undrawn (DR 1)

filaments is comparable to viscose (340 MPa, Adusumali et al. (2006)). A

function of the form σ = σmax(1–a/DR) was employed to describe the development

of tenacity vs. DR. If no extrusion velocity dependency is assumed, the parameters

are σmax = 39.66±0.87 cN tex–1 and a = 0.51±0.04. Indeed, there was no trend with

extrusion velocity within the tested range. Although the scatter was too high to

evaluate any correlation, a 3rd degree polynomial surface is added in Fig. 4 assist

with visualization of the data points. Elongation at break was 14.6±1.7% at DR 1,

but already at DR 3, it reduced to 9.1±1.4%, i.e. statistically consistent with its

ultimate value of 8.2±1.8% for DR ≥ 5 (Fig. S2). This is significantly smaller than

values obtained for other ILs (11.2–15.5%), but comparable to Lyocell (Kosan et

al. 2008; Adusumali et al. 2006).

Page 12: Dry jet-wet spinning of strong cellulose filaments from

11

Fig. 4 Tenacity (cN tex–1) vs. draw ratio and extrusion velocity (ml min–1); the vertical scale is

inverted for clarity

The elastic modulus (Fig. 5) was rather high, up to 15.4±2.0 cN tex–1 %–1

(23.2±3.1 GPa), which is in the range of Lyocell and rayon tirecord fibers

(Adusumali et al. 2006). The maximum was reached at DR 5. However, it returned

to 11.9±2.2 cN tex–1 %–1 (17.8±3.3 GPa) at DR 12.5. This, along with the absence

of changes in other parameters suggests that at DR 7.5 and beyond, the filament is

probably slipping on the godet, causing fluctuations of the instantaneous DR, thus

causing damage (slight lateral cracks). This is supported by the observation that

the filament must be still wet when stretched. If the stretching was done while the

filaments were dry, this would be obviously apparent in filament testing: the

elastic region would be absent and only plastic deformation would occur with a

modulus of ca. 5 cN tex–1 %–1, a behavior which is not observed (Gindl et al.

2008b).

The proportional limit was 1.05±0.09%, in a manner independent of DR and ve

(Fig. S3). A durable and processable yarn would have a high elastic resilience, i.e.

the strain energy density where the material is yet undamaged. Indeed, resilience

(Fig. S4) was found to be largely constant at 1.10±0.32 MJ m–3, which is very

similar to Lyocell (0.8–1.6 MJ m–3) (Gindl et al. 2008a). DR 12.5 filaments had a

higher limit of 1.25±0.24% and a correspondingly higher resilience of 1.39±0.78

MJ m–3, but due to the scatter—again probably because of damage—this is not

statistically significant.

Page 13: Dry jet-wet spinning of strong cellulose filaments from

12

Fig. 5 Elastic modulus vs. draw ratio and extrusion velocity (ml min–1)

Orientation

The total orientation starts at a high level (58±3%) and immediately increases to

the final value (69±5%), and is not a function of extrusion velocity: birefringence

is described by the function Δn = (0.044±0.001) – (0.0080±0.0023)/DR, with no

pattern in the residuals (see Figs. 6, S5). Qualitatively, this is in agreement with

results with NMMO by Mortimer et al., who found Δn starting at a constant value

(Δn = 0.023) and tending towards an asymptote (0.041, Mortimer et al. (1996)).

Quantitatively, our filaments start even higher (0.036±0.002) although the final

value is similar (0.043±0.003).

Northolt proposed a correlation between birefringence and compliance (E–1, the

inverse of modulus) (Northolt 1985; Kong and Eichhorn 2005). With our data, the

correlation gave the shear modulus of g = 0.60±0.06 GPa (see Fig S6), a rather

low value if compared to g = 2.5 GPa from Northolt (1985) or 3.6 GPa from Kong

and Eichhorn (2005). However, the fit was not good, because low compliances are

not reached, probably because of the fibers being damaged at high draw ratios.

Additionally, tenacity was not significantly correlated with birefringence for

DR>1.

Page 14: Dry jet-wet spinning of strong cellulose filaments from

13

Fig. 6 Birefringence vs. draw ratio and extrusion velocity (ml min–1)

Wet ultimate tensile strength was 76±28% of dry strength for DR>1, and 67±18%

for DR 1, unaffected by ve (Fig S7). Similarly, elongation at break when wet was

143±49% of dry strength for DR>1, and 137±55% for DR 1 (Fig S8). This is

concurrent with the trends in total orientation (constant at DR>1). The explanation

for the independence of the ratio from DR and ve is that the orientations of

crystalline and amorphous parts increase in lockstep, because the deformation is

so fast that there is no time for relaxation and crosstalk between the proto-

crystalline and proto-amorphous domains in the jet (Fig S9). Or, by

contraposition, a simple increase in crystallinity does not cause the increase in

tenacities, nor do the crystallites orient independently without concomitant

amorphous orientation. In contrast to Jiang et al. (2012), speed had no observable

effect, since for our very high-viscosity dope, the viscous contribution to the line

stress dominates over other contributions such as inertial and drag. The same view

is also supported by a similar argument with wet modulus. The wet modulus is

traditionally reported as the BISFA modulus, which is the stress at 5% strain

while immersed in water (BISFA 2004). Since wet fibers have a simple, almost

exactly linear stress-strain curve ending in a random break, this is a valid proxy

for the true wet modulus. In Fig. 7, it can be seen that the BISFA modulus follows

a rather different pattern from the dry modulus, with a near-linear increase to a

maximum at DR 8. The ratio of dry to wet modulus (Fig. 8) decreases from

9.2±0.5 to 4.2±0.6 from DR 1 to 10, which is in quantitative agreement with

results from NMMO (Coulsey and Smith 1996). Namely, there is a smooth

Page 15: Dry jet-wet spinning of strong cellulose filaments from

14

transition from randomly oriented amorphous "defects" aligning into an axial

orientation, transferring the load from the swollen defects into aligned crystallites

and thus significantly increasing the stiffness when wet. Thus, the key parameter

for mechanical properties is suggested to be the orientation of amorphous domains

rather than aggregate quantities such as total orientation or crystallinity. An

unbroken load path through crystallites (Fig. S9) gives the filament its high

modulus. The low absolute modulus at high DR is incidental and caused by

slipping, and can be corrected by technical improvements in the process.

Orientation inferred from Fig. 8 continues to improve to ca. DR 5, which

demonstrates that the system behaves according to the Kratky II limiting case

(linked chains drawn axially into alignment) rather than the Kratky I limiting case

(rigid rods being forced together) (Ziabicki 1976; Mortimer et al. 1996). However,

tenacities improve only slightly after DR 5 and cannot continue to improve with

draw if lateral defects remain or are created, because even a single significant

defect along the 20 mm measurement length can cause early rupture in the tensile

strength test. The damage to the filament can be inferred from the modulus data

(Fig. 5), because filaments are stretchier at weak points due to a smaller effective

area, reacting with greater deformation to the same force.

Fig. 7 BISFA modulus vs. draw ratio and extrusion velocity (ml min–1)

Page 16: Dry jet-wet spinning of strong cellulose filaments from

15

Fig. 8 Ratio of dry to wet modulus with an exponential fit

Effects of the aspect ratio of the spinneret and guide-to-godet stress

A spinneret has a convergent conical exit nozzle, terminating in a short straight

capillary, and the aspect ratio is the length-to-diameter ratio (L/D) of the latter.

For comparison, filament was spun with a L/D 0.2 spinneret and ve = 0.04 ml

min–1 and compared to L/D 2.0 results above (Figs 3, 4, 5). First, the maximum

draw ratio was 5.0 instead of 7.5, with consistent breaks at DR 5.5 and 7.5.

Furthermore, there were no statistically significant differences in titer, modulus or

tenacity (Figs 2 and 9), and thus the correction factor s remained the same at 2

(exactly, s = 1.95±0.08). For NMMO, a slight birefringence of the jet near the

nozzle has been observed (Mortimer and Péguy 1996a), and better strength

properties at a given DR can be obtained by increasing L/D (Zikeli et al. 1992),

which suggests that a longer spinneret gives a slightly better orientation. Indeed,

at DR 5, total orientation at L/D 0.2 was 0.65±0.11 or slightly lower than for L/D 2

(0.72±0.11), albeit the difference is not statistically significant.

Given that breaks occur around the second guide, the question is if reducing the

stress there would give access to higher draw ratios. Thus, the second guide was

connected to the godet with a rubber belt, giving it equal linear speed. The DR 1

filament was indeed thinner, but the DR 5 filament was unaffected (Figs 2 and 9).

The breaks occur in the bath, and the same limit remains; thus, the limit may be

Page 17: Dry jet-wet spinning of strong cellulose filaments from

16

because of stress on the partially regenerated filament in and immediately after

the bath rather than only because of the stress at the observed breakpoint.

Fig. 9 Effects of aspect ratio and guide rotation on tenacity and modulus; some tenacity points

offset on x-axis for clarity

Effect of bath temperature

Given that diffusion is faster at higher temperatures, it is natural to ask if the bath

temperature could be increased. However, relaxation of the polymer chains is also

promoted by higher temperatures (Ziabicki 1976). To resolve this question,

coagulation bath temperatures 15, 30 and 45 ºC were explored (with ve = 0.04 ml

min–1, L/D 2, Table 1). The incipient filament was weakened: maximum DR was

limited to 3 at 30 ºC and to 1 at 45 ºC. In the Lyocell process, cooling in the air

gap effectively solidifies the jet before regeneration (Mortimer et al. 1996), giving

access to high draw ratios. Reheating in the bath halts this process, allowing for

more polymer relaxation and causing a loss of orientation. As a result, with the

bath at 30–45 ºC, elongation at break was higher and modulus and resilience were

lower. For drawn (DR 3) filaments the tenacity (24 cN tex–1) was similar to

viscose. A decrease in orientation was apparent from the drop in birefringence:

total orientation decreased from 57±2% to 42±5% for DR 1 and from 70±3% to

59±3% for DR 3. Thus, proper solidification of the filament prior to regeneration

is crucial: in a warmer bath, orientation is lost, producing a relatively amorphous

structure that can deform plastically but has low tensile strength.

Table 1 Effects of increasing bath temperature on filament properties

Page 18: Dry jet-wet spinning of strong cellulose filaments from

17

DR bath

[ºC]

titer

[dtex]

tenacity

[cN tex–1]

elongation

[%]

modulus

[cN tex–1 %–1]

resilience

[MJ m–3]

birefringence

1 15 24.0±3.0 18.9±2.0 16.3±2.8 6.9±1.0 0.9±0.3 0.036±0.001

30 22.8±3.5 16.4±2.3 28.3±8.5 5.7±1.1 0.7±0.3 0.026±0.002

45 24.9±3.6 14.0±1.8 29.0±5.4 4.7±0.6 0.7±0.2 0.026±0.003

3 15 7.9±1.2 35.1±4.4 9.7±1.2 13.1±1.7 1.0±0.2 0.043±0.002

30 7.6±0.5 24.1±1.7 12.6±1.4 10.1±1.3 0.9±0.2 0.037±0.002

7.5 15 3.0±0.9 38.5±8.4 7.9±1.7 14.5±1.0 1.2±0.2 0.042±0.002

Conclusions

A dry-jet wet spinning process for spinning cellulose fibers from an ionic liquid

solution is characterized and its limitations and sensitivities discussed. A 13 wt-%

solution of dissolving pulp in an ionic liquid ([DBNH][OAc]) was extruded and

drawn in air and regenerated in water. With extrusion velocities in the range 0.02

to 0.04 ml min–1, drawing in the air gap with DR 7.5–12.5 gave filaments with

high tenacities (36.8±2.2 cN tex–1 or 552±34 MPa) and moduli (15.4±2.0 cN tex–1

%–1 or 23.2±3.1 GPa). Linear density of the finished fibers was proportional to

DR–1. This suggests a simple deformation, rather than the more complicated

deformation seen in NMMO spinning, where the viscosity increase in the air gap

due to cooling leads to a dependency on DR–0.5 instead (Kong and Eichhorn 2005).

Mechanical properties (tenacity, modulus, resilience) and birefringence were

independent of extrusion velocity, and orientation was nearly complete at DR 5.

The lack of a speed dependency suggests again a simple deformation, where

amorphous domains are stretched and thus oriented in lockstep with crystalline

regions. This is explicitly revealed by dry/wet modulus ratio changing from 9 to 4

with increasing draw ratio. Strong filaments were readily spun only with a

coagulation bath temperature of 15 °C, but not with 30 or 45 °C. A spinneret

aspect ratio (L/D) of 2.0 gave a better orientation than L/D 0.2 and allowed to

reach higher draw ratios. However, our small-scale monofilament setup seems to

slightly damage high-draw ratio filaments, because there is slipping on the takeup

godet. Thus, improvement is expected when spinning in multifilament mode,

where the fiber bundle is stronger and less likely to slip than a single filament. In

summary, the process was found to be robust and simple even with the suboptimal

setup of a small monofilament system, and in performance comparable to the

commercial NMMO-based Lyocell process.

Page 19: Dry jet-wet spinning of strong cellulose filaments from

18

Acknowledgements The authors wish to thank TEKES – the Finnish Funding Agency for

Technology and Innovation and Finnish Bioeconomy Cluster FIBIC Ltd. for funding in the

framework of the Future Biorefinery Cellulose project. Arno Parviainen synthesized the

[DBNH][OAc] with Dr. Alistair King and Prof. Ilkka Kilpeläinen (University of Helsinki). Dope

samples were kindly provided by Shirin Asaadi. Kaarlo Nieminen assisted with MatLab and GPC

was operated by Lasse Tolonen.

References

Adusumali R-B, Reifferscheid M, Weber H, Roeder T, Sixta H, Gindl W (2006) MechanicalProperties of Regenerated Cellulose Fibres for Composites. Macromol Symp 244 (1):119-125. doi:10.1002/masy.200651211

ASTM (2010) Standard Test Method for Tensile Properties of Yarns by the Single-Strand Method.D2256/D2256M. ASTM International, West Conshohocken, PA, United States

Bernot RJ, Brueseke MA, Evans-White MA, Lamberti GA (2005) Acute and chronic toxicity ofimidazolium-based ionic liquids on Daphnia magna. Environ Toxicol Chem 24 (1):87-92.doi:10.1897/03-635.1

BISFA (2004) Testing methods viscose, modal, lyocell and acetate staple fibres and tows. BISFA -The International Bureau for the Standardisation of Man-Made Fibres, Brussels

Coulsey HA, Smith SB (1996) The formation and structure of a new cellulosic fibre. LenzingerBer 75:51-61

Cross MM (1965) Rheology of non-Newtonian fluids: A new flow equation for pseudoplasticsystems. Journal of Colloid Science 20 (5):417-437. doi:10.1016/0095-8522(65)90022-X

Doherty TV, Mora-Pale M, Foley SE, Linhardt RJ, Dordick JS (2010) Ionic liquid solventproperties as predictors of lignocellulose pretreatment efficacy. Green Chem 12(11):1967. doi:10.1039/c0gc00206b

Fink HP, Weigel P, Purz HJ, Ganster J (2001) Structure formation of regenerated cellulosematerials from NMMO-solutions. Prog Polym Sci 26 (9):1473-1524. doi:10.1016/s0079-6700(01)00025-9

Gindl W, Reifferscheid M, Adusumalli RB, Weber H, Röder T, Sixta H, Schöberl T (2008a)Anisotropy of the modulus of elasticity in regenerated cellulose fibres related tomolecular orientation. Polymer 49 (3):792-799. doi:10.1016/j.polymer.2007.12.016

Gindl W, Reifferscheid M, Martinschitz KJ, Boesecke P, Keckes J (2008b) Reorientation ofcrystalline and noncrystalline regions in regenerated cellulose fibers and films tested inuniaxial tension. Journal of Polymer Science Part B: Polymer Physics 46 (3):297-304.doi:10.1002/polb.21367

Grasa GA, Kissling RM, Nolan SP (2002) N-Heterocyclic Carbenes as Versatile NucleophilicCatalysts for Transesterification/Acylation Reactions. Org Lett 4 (21):3583-3586.doi:10.1021/ol0264760

Gries T, Wirth B, Warnecke M, Schmenk B, Seide G (2011) Filament breaches during air-gapspinning. Chemical Fibers International 61:38-39

Happel S (2014) Lenzing Fibers. Paper presented at the Mistra Future Fashion Symposium,Stockholm, Sweden, 27 May 2014

Hauru LKJ, Hummel M, King AWT, Kilpeläinen IA, Sixta H (2012a) Role of solvent parametersin the regeneration of cellulose from ionic liquid solutions. Biomacromolecules 13(9):2896-2905. doi:10.1021/bm300912y

Hauru LKJ, Hummel M, Sixta H (2012) Fibre spinning from ionic liquid dope. In: 12th EuropeanWorkshop on Lignocellulosics and Pulp, Espoo, Finland, August 27-30, 2012. Universityof Helsinki, Helsinki, pp 272-275

Hauru LKJ, Ma Y, Hummel M, Alekhina M, King AWT, Kilpeläinen IA, Penttilä PA, Serimaa R,Sixta H (2013) Enhancement of ionic liquid-aided fractionation of birchwood. Part 1:Autohydrolysis pretreatment. RSC Advances 3:16365-16373. doi:10.1039/C3RA41529E

Hearle JWS, Schawaller D (2000) Fibers, 2. Structure. In: Ullmann's Encyclopedia of IndustrialChemistry. Wiley-VCH Verlag GmbH & Co. KGaA.doi:10.1002/14356007.a10_477.pub3

Hummel M, Michud A, Sixta H Extensional rheology of cellulose-ionic liquid solutions. In:Nordic Rheology Conference, June 8-10, 2011 2011. p 313

Page 20: Dry jet-wet spinning of strong cellulose filaments from

19

Hummel M, Michud A, Tanttu M, Asaadi S, Ma Y, Hauru LKJ, Parviainen A, King AWT,Kilpeläinen I, Sixta H (2014) Ionic liquids for the production of man-made cellulosicfibers- opportunities and challenges. Adv Polym Sci submitted

Hämmerle FM (2011) The cellulose gap (the future of cellulose fibers). Lenzinger Ber 89:12-21Jiang G, Yuan Y, Wang B, Yin X, Mukuze K, Huang W, Zhang Y, Wang H (2012) Analysis of

regenerated cellulose fibers with ionic liquids as a solvent as spinning speed is increased.Cellulose 19 (4):1075-1083. doi:10.1007/s10570-012-9716-2

King AWT, Asikkala J, Mutikainen I, Järvi P, Kilpeläinen I (2011) Distillable Acid-BaseConjugate Ionic Liquids for Cellulose Dissolution and Processing. Angew Chem 123(28):6425-6429. doi:10.1002/ange.201100274

King AWT, Parviainen A, Karhunen P, Matikainen J, Hauru LKJ, Sixta H, Kilpelainen I (2012)Relative and inherent reactivities of imidazolium-based ionic liquids: the implications forlignocellulose processing applications. RSC Advances 2:8020-8026.doi:10.1039/c2ra21287k

Kong K, Eichhorn SJ (2005) Crystalline and amorphous deformation of process-controlledcellulose-II fibres. Polymer 46 (17):6380-6390. doi:10.1016/j.polymer.2005.04.096

Kosan B, Michels C, Meister F (2008) Dissolution and forming of cellulose with ionic liquids.Cellulose 15 (1):59-66. doi:10.1007/s10570-007-9160-x

Krässig HA (1993) Cellulose: structure, accessibility and reactivity, vol 11. Polymer Monographs.Gordon and Breach Science Publishers, Chemin de la Sallaz, Switzerland

Laus G, Bentivoglio G, Schottenberger H, Kahlenberg V, Kopacka H, Röder T, Sixta H (2005)Ionic liquids: current developments, potential and drawbacks for industrial applications.Lenzinger Ber 84:71-85

Lenz J, Schurz J, Wrentschur E (1994) On the Elongation Mechanism of Regenerated CelluloseFibres. Holzforschung 48 (s1):3-158. doi:10.1515/hfsg.1994.48.s1.72

Michud A, King A, Parviainen A, Sixta H, Hauru L, Hummel M, Kilpeläinen I (2014) Process forthe production of shaped cellulose articles from a solution containing pulp dissolved indistillable ionic liquids. FI Patent Application PCT/FI2014/050238, 2014-04-04

Mortimer SA, Péguy AA (1996a) The formation of structure in the spinning and coagulation ofLyocell fibres. Cellul Chem Technol 30:117-132

Mortimer SA, Péguy AA (1996b) The influence of air-gap conditions on the structure formation oflyocell fibers. J Appl Polym Sci 60 (10):1747-1756. doi:10.1002/(SICI)1097-4628(19960606)60:10<1747::AID-APP28>3.0.CO;2-#

Mortimer SA, Péguy AA, Ball RC (1996) Influence of the physical process parameters on thestructure formation of Lyocell fibres. Cellul Chem Technol 30:251-266

Northolt MG (1985) Are Stronger Cellulose Fibres Feasible? Lenzinger Ber 59:71-78Parviainen A, King AWT, Mutikainen I, Hummel M, Selg C, Hauru LKJ, Sixta H, Kilpeläinen I

(2013) Predicting Cellulose Solvating Capabilities of Acid–Base Conjugate IonicLiquids. ChemSusChem 6 (11):2161-2169. doi:10.1002/cssc.201300143

Röder T, Moosbauer J, Kliba G, Schlader S, Zuckerstätter G, Sixta H (2009) Comparativecharacterisation of man-made regenerated cellulose fibers. Lenzinger Ber 87:98-105

Schrems M, Ebner G, Liebner F, Becker E, Potthast A, Rosenau T (2010) Side Reactions in theSystem Cellulose/1-Alkyl-3-methyl-imidazolium Ionic Liquid. In: Cellulose Solvents:For Analysis, Shaping and Chemical Modification, vol 1033. ACS Symposium Series,vol 1033. American Chemical Society, pp 149-164. doi:10.1021/bk-2010-1033.ch008

Schuster KC, Aldred P, Villa M, Baron M, Loidl R, Biganska O, Patlazhan S, Navard P, Rüf H,Jericha E (2003) Characterising the emerging Lyocell fibres structures by ultra smallneutron angle scattering (USANS). Lenzinger Ber 82:107-117

Stibal W, Schwarz R, Kemp U, Bender K, Weger F, Stein M (2000) Fibers, 3. General ProductionTechnology. In: Ullmann's Encyclopedia of Industrial Chemistry. Wiley-VCH VerlagGmbH & Co. KGaA. doi:10.1002/14356007.a10_511

Sun N, Swatloski RP, Maxim ML, Rahman M, Harland AG, Haque A, Spear SK, Daly DT,Rogers RD (2008) Magnetite-embedded cellulose fibers prepared from ionic liquid. JMater Chem 18 (3):283-290. doi:10.1039/b713194a

Swatloski RP, Spear SK, Holbrey JD, Rogers RD (2002) Dissolution of Cellose with IonicLiquids. J Am Chem Soc 124 (18):4974-4975. doi:10.1021/ja025790m

Viswanathan G, Murugesan S, Pushparaj V, Nalamasu O, Ajayan PM, Linhardt RJ (2006)Preparation of Biopolymer Fibers by Electrospinning from Room Temperature IonicLiquids. Biomacromolecules 7 (2):415-418. doi:10.1021/bm050837s

Zhang H, Wang ZG, Zhang ZN, Wu J, Zhang J, He JS (2007) Regenerated-Cellulose/Multiwalled-Carbon-Nanotube Composite Fibers with Enhanced Mechanical Properties Prepared with

Page 21: Dry jet-wet spinning of strong cellulose filaments from

20

the Ionic Liquid 1-Allyl-3-methylimidazolium Chloride. Adv Mater (Weinheim, Ger) 19(5):698-704. doi:10.1002/adma.200600442

Ziabicki A (1976) Fundamentals of Fibre Formation. John Wiley & Sons, LondonZikeli S, Firgo H, Eichinger D, Jurkovic R (1992) Verfahren zur Herstellung eines cellulosischen

Formkorpers. European Patent Application EP 0 494 852 A2, 1992-07-15