• No results found

Optical Coherence Tomography measurements of biofibril dispersions in flow-focusing

N/A
N/A
Protected

Academic year: 2021

Share "Optical Coherence Tomography measurements of biofibril dispersions in flow-focusing"

Copied!
46
0
0

Loading.... (view fulltext now)

Full text

(1)
(2)
(3)

Optical Coherence Tomography measurements of biofibril dispersions in flow-focusing

Cecilia Rydefalk

(4)
(5)

Contents

1 Introduction 4

1.1 Nano-fibrils . . . 4

1.2 Assembly of fibrils into filaments . . . 4

2 Method 8 2.1 Experimental setup . . . 8

2.2 Optical Coherence Tomography . . . 8

2.3 Post-processing . . . 12

3 Measurements on cellulose nano-fibril dispersions 14 3.1 Material . . . 14

3.2 Flow-focusing geometries and parameters . . . 15

3.3 Results . . . 16

3.4 Discussion . . . 19

4 Measurements on protein nano-fibril dispersions 22 4.1 Material . . . 22

4.2 Flow-focusing geometries and parameters . . . 23

4.3 Results and observations . . . 24

4.4 Discussion . . . 29

5 Artefacts and errors 30

6 Concluding remarks 32

A Secondary flow patterns 33

(6)
(7)

Abstract

Microfluidic spinning techniques have been used as a successful assembly process for biofibrils dispersed in a fluid. The dispersion flows through a microchannel and is focused by a sheath flow. In this way the topology and velocity of the dispersion in the core can be manipulated. By using Optical Coherence Tomography this topological development and velocities in the channel can be investigated and vi- sualized. This is done for two different biofibrils: cellulose nano-fibrils and protein nano-fibrils. In both cases the geometry of the channel has been varied to investi- gate how this affects the flow of the dispersion. In the case of cellulose nano-fibrils the inlet angle of the side channels in the focus-section is varied. There is a thread forming for all angles tested, but with variations in shape. The velocity measure- ments also show differences in acceleration and deceleration behaviour; indicating differences in inner morphology. For the protein nano-fibrils the distance between the side-channel inlets, in a device with two focus-sections, is shown to affect the wetting behaviour of the dispersion.

Sammanfattning

Att anv¨anda mikrostr¨omning f¨or att spinna fibrer av biofibriller har visas sig vara en framg˚angsrik metod. En dispersion best˚aende av av en v¨atska med fibrillerna i f˚ar str¨omma genom en mikrokanal med en str¨omningsriktande sektion. D¨ar in- troduceras ett fl¨ode fr˚an sidorna som anv¨ands f¨or att manipulera topologi och hastighet hos dispersionen. Genom att anv¨anda optisk koherenstomografi kan den topologiska utvecklingen och hastigheten hos dispersionen studeras och vi- sualiseras. Det har h¨ar gjorts f¨or tv˚a olika biofibriller: cellulosa-nanofibriller och protein-nanofibriller. I b˚ada fallen har kanalens geometri varierats f¨or att kunna j¨amf¨ora hur detta inverkar p˚a str¨omningen. F¨or cellulosa-nanofibrillerna varierades inloppsvinkeln p˚a sidokanalerna i den str¨omningsriktande sektionen.

En dispersionstr˚ad bildas f¨or alla geometrifall, men med variationer i formen.

Hastighetsm¨atningarna visar skillnader i acceleration och retardation vilket in- dikerar skillnander ¨aven in tr˚adarnas inre struktur. F¨or protein-nanofibrillerna varierades avst˚andet mellan de str¨omningsriktande sektionerna, i en kanal med tv˚a s˚adana. Detta visar sig p˚averka hur dispersionstr˚aden sl¨apper fr˚an kanalv¨aggarna.

(8)
(9)

Acknowledgements

The completion of this thesis was not made according to the cookbook saying that the thesis is written at the end of a four year period of research and theoretical studies [30].

Thus, starts the foreword of my father’s doctoral thesis1, but it might as well have been my own words regarding my studies at KTH. For reasons I won’t go into, my time at KTH has stretched over a period of over 15 years, and taken different forms, and is finally culminating in this thesis. There are therefore plenty of people to whom I owe thanks.

My advisor, Fredrik Lundell, for letting me do this project, for enthusiasm and encouragement along the way, and for letting me be a part of this amazing group of extraordinary people he has gathered around himself. It takes a village to raise a child, or in this case it took a research group to raise this student.

I am forever grateful to everyone who contributed with their knowledge, time, patience, and moral support. First and foremost, Krishne who’s been the day-to- day advisor. Sagar, Martin, Korneliya, Rena, Jakob, Gusten, Susumu and Calvin, who all contributed in some way at some point. I’ve had so much fun and I’ve learned much outside the scope of my thesis that I will take with me in all my future adventures. I can no other answer make but thanks, and thanks, and ever thanks; and oft good turns are shuffled off with such uncurrent pay.

The guys I’ve studied with during the master’s program. You’ve made my life both harder and easier in all the best ways. Mostly you’ve been awesome day-to- day support. I hope to keep seeing you in the future even as we are all done now.

Since my dear soul was mistress of her choice, and could of men distinguish, her election hath sealed thee for herself.

My dearest friends whom I’ve barely seen these past few years while pursuing this long-standing dream of mine to finally become an engineer, but who have assured me that I’m “worth waiting for”. I look forward to finally seeing you again. I count myself in nothing else so happy as in a soul remembering my good friends.

My family that have seen very little of me during all this, but have supported me all the way. And Kuma, with you everything becomes possible. I would not wish any companion in the world but you.

1A work that also concerned particle dispersions.

(10)
(11)

Chapter 1 Introduction

1.1 Nano-fibrils

Nature creates the most marvellous materials from small and rather simple building blocks, like nano-fibrils. These are nano-sized, slender biopolymers that are found in living organisms. By stacking them in hierarchical structures the properties of the fibrils, e.g. strength and stiffness, can be transferred to macro-scale materials, such as spider silk, wood and exoskeletons of crustaceans. These are materials that are very strong in comparison to their density [24]. Efforts have been made to mimic this in engineered materials, and utilize the qualities of the nano-fibrils to build macro materials [4, 6, 25]. However, it is difficult to assemble the nano-fibrils with precision. Misalignment, poor adhesion and defects on nano-level appearing at the macro-level, due to the assembly processes, makes for a weaker material.

[24, 25]. Even so, the last few years have seen plenty of progress in this field with the creation of films, foams and filaments from nano-particles, with possible applications in a wide variety of fields: from paper-batteries to ways of treating burn victims [16, 29]. The fibrils have a typical size of 2 − 5 nm in diameter and up to 1µm in length [22, 24, 26]. This work is focused on understanding the flow dynamics of the filament-spinning process.

1.2 Assembly of fibrils into filaments

One way of assembling nano-fibrils into a macro-material is to disperse them in a fluid, and then use hydrodynamic alignment to form filaments; a process also known as micro-fluidic spinning. Figure 1.1 shows the principle of a double flow- focusing channel used for spinning. The dispersion flows in from the left. In the first cross-shaped flow-focusing section a sheath fluid enters, thereby accelerating the flow and aligning the particles. In the second a gelling agent is introduced to

(12)

Figure 1.1: Principle of the double flow-focusing channel.

further increase the alignment and lock the fibril structure in the filament.

From a material manufacturing point of view, there are four key aspects to consider when making a filament. First, the detachment of the nano-fibril disper- sion from the top and bottom walls, and the formation of a stable thread. Second, the aspect ratio of the thread. Third, the velocity changes along the channel. And fourth, when to introduce the gelling agent. All of these aspects will, in the end, affect the material properties of the final filament.

Breaking down the process of filament spinning into the physics of the flow dynamics, there is still much that needs to be understood about this system.

Similar micro-scale flow-focusing systems are utilized for the formation of droplets and bubbles, and some can be learnt from this area, such as regimes in which two immiscible fluids form different structures [2, 7, 8]. Different flow regimes, depending on viscosity, velocity and interfacial tension, have been mapped out into into structural categories such as jetting, tubing, dripping, engulfment and threading [7]. A thread, in this context, is an extruded fluid stream with a core moving in a solid-like motion without contact with the walls [8]. This means that somewhere around 2 in Figure 1.2 the core fluid detaches from the upper and lower wall and becomes completely surrounded by the sheath1. The transition between the flow regimes depends mainly on the capillary numbers of the fluids involved. The capillary number Ca of a fluid flow is defined as:

Ca = ηU

γ (1.1)

where η is the viscosity, U is the velocity and γ the interfacial tension between the fluids.

1The regime where the core stays in contact with the upper and lower walls and only has sheath fluid along the side walls is called tubing [7].

(13)

Figure 1.2: The side-view at the center of the channel, and the view from above at the center of the channel, shown above the changes in velocity along the centerline of the colloid.

For hydrodynamic spinning of nano-fibril filaments, the threading regime is of interest. When spinning nano-fibril dispersions, however, the fluids involved are miscible. This means that the interface between the core and sheath fluids is transient and eventually full mixing will occur [10]. For immiscible fluids there is an interfacial tension between the two fluids, and under certain conditions there is something called effective interfacial tension, γ, also between miscible fluids. The effective interfacial tension is given as

γ = K∆Φ2

δ (1.2)

where K is the Korteweg constant of the system, δ is the thickness of the interface, which is governed by the diffusion coefficient, and ∆Φ is the change in concentra- tion, or volume fraction Φ [10, 28, 34, 40]. The different configurations of viscosity ratios, velocity ratios, interfacial tension and geometries will then impact if we are in the threading regime and the aspect ratio achieved.

The flow also controls the nano-structure inside the thread. For stiff nano-rods experiments have shown how velocity variations in the micro-channel affects the

(14)

order parameter2 of the fibrils [5]. The way the fibrils are ordered will, in turn, affect e.g. mechanical properties and electrical conductivity of the final material [5, 13, 17]. Figure 1.2 shows the principle of the thread formation, and the velocity change in the center of the core, in the flow-focusing device. The core fluid, the dispersion, enters from the left with pump flow rate Q1. In the focusing section the sheath fluid enters with pump flow rates Q2. The sheath enters at 1 , and the velocity decrease, the fibrils tend to align perpendicular with respect to the flow direction. The flow then accelerates at 2 when it enters the downstream part of the channel after the focusing. Further downstream 3 the velocity profile is flat across the core.

It has been shown that long and short fibrils respond differently to the velocity variations, meaning that for a polydisperse system there are possibilities of different configurations of long and short fibrils, e.g. aligned long ones with cross-linked short fibrils perpendicular to them [5].

After aligning the fibrils, at 2 in Figure 1.2, Browninan motion will drive them towards isotropy again. Therefore, the last step in material processing is chemically locking the structure by introducing a gelling agent while they are still in the desired configuration. Gelling is not studied in the present work.

Purpose of the studies

There are two separate studies presented in this thesis. The first one concerns cellulose nano-fibrils. The parametric study performed gives insight into the pos- sibilities to further tailor materials and also provides data for validation of a simu- lation model that explores the effects of effective interfacial tension and geometry variations. The second study deals with protein nano-fibrils. The purpose here is to visualize the thread under specific micro-fluidic spinning conditions, varying the channel geometry and the morphology of the protein nano-fibrils, providing knowledge necessary for future process development.

2A characterization of the alignment of the particles of the system where 0 is completely random orientation and 1 is complete alignment with the director and -0.5 is anti-alignment or perpendicularity against the director [18]. Here the director is the flow direction.

(15)

Chapter 2 Method

2.1 Experimental setup

The micro-channels used in these experiments have been made by electrical dis- charge machining of 1 mm wide slits through a 1 mm thick stainless steel plate, thus creating a square channel when the plate is covered with a sheet of 140µm thick Cyclic Olefin Copolymer (COC) film on each side, through which optical measurements can be made. This is sandwiched between two aluminum plates with milled slits for optical access, see Figure 2.1a. The inlets are connected to syringe pumps (WPI AI-4000). The principle of the channel layout is shown in Figure 2.1d with the corresponding coordinate system employed in this thesis in Figure 2.1b. The core, in blue, entering with Q1 is surrounded by the sheaths, Q2, from the side channels. In the experiments either the inflow angle of the flow-focus, β, is varied or the distance between a pair of flow-focus sections (see Figure 1.1 in the previous chapter). For devices, with several flow-focus sections, the volumetric flow rates are numbered in descending order from upstream to downstream (Q2 and Q3). For reasons described in the Section 2.2, the channel is tilted at an angle α with regard to the horizontal, so as not to be perpendicular to the beam in the measuring device. This means that the cross-section of the channel seen by the measurement device is at a slant with regard to the flow direction. This slanted section is referred to as the cross-section throughout the thesis.

2.2 Optical Coherence Tomography

The method used to study the flow in the channel is Optical Coherence Tomogra- phy (OCT). This is a non-intrusive measuring technique that can provide high res- olution 3D-images from optical scattering media. By using Phase-resolved Doppler OCT the velocity field can be acquired, in addition to the structural information

(16)

(a) (b) (c)

(d)

Figure 2.1: (a) The sandwich structure holding the channel. 1 are the aluminum plates, 2 the plastic film and 3 the metal sheet with the channel. (b) Coordinate system employed in the thesis. (c) The cross-section of the (downstream) channel showing the non-dimensionalized height and width of the thread, εz/h and εx/h.

(d) Schematic of the micro-channel with the flow-focusing section, with the core flow in blue.

(17)

of the flow.

Other methods for visualizing and studying the structure of the flow in a flow- focusing device of this scale is, for example, to use transparent channels and high- speed cameras together with a high magnification lens [8] or a microscope [17], but this will limit the study to 2D images. To measure the velocities in the microchannel it is possible to use µPIV, a velocimetry technique that identifies and tracks individual particles in the flow. Additional particles need to be mixed with the dispersion for the measurements [12].

The advantage of using OCT is its short measuring time, that it is easy to handle and will provide both high resolution 3D-images of the phases and flow inside of the channel. The dispersions in this study did not require any seeding of additional particles, thereby conserving the physiochemical properties of the core.

(a) (b)

Figure 2.2: (a) The principle of the interferometry of the OCT as described in the Telesto manual [33]. (b) The build up of A-scans into 2D- and 3D-images.

Principle of OCT

Optical Coherence Tomography (OCT) is based on the Michelson interferometer.

It uses low-coherent light, and the light source is split into two beams that are compared to each other when reflected back. One as a reference, the other reflected on the media one wishes to study [32,33]. This is done through the depth of turbid media returning information of the amount of back-scattered light over depth in form of a dB-curve. One depth scan in a line is called an A-scan. Consecutive A-scans can be performed generating a B-scan which gives a 2D-image of the

(18)

cross-section of the media. Several consecutive B-scans will then form a grid of 3D-data from the sample, see Figure 2.2. This is translated to intensity of the back-scattered light, and if the medias studied have sufficiently different intensity:

they can be distinguished from each other in the images.

However, if there are intense reflections from the sample, they can cause ad- ditional harmonic frequencies to appear risking saturation of the images. In this work the micro-channel it is tilted with respect to the A-scan axis to reduce this effect [33].

Phase-resolved Doppler OCT

In addition to measuring the intensity of the back-scattered light, a Phase-resolved Doppler-OCT also measures the phase shift. For measurements of flowing media the phase-shift between successive A-lines at the same depth can be used to acquire the magnitude and direction of the axial velocity (parallel to the imaging beam) [35] as shown in Figure 2.3. This means that the channel needs to be at an angle with respect to the beam so that the the main flow direction and the beam are not perpendicular to each other as this velocity will be vs = v||/ cos θ, where θ is the angle between the beam and the flow direction. The velocity from the phase-shift

∆φ can then be retrieved as:

vs= ∆φ λf

4nπ cos θ = ∆φ λf

4nπ sin α (2.1)

where θ is the angle between flow direction and beam, α is the angle of the channel with regard to the horizontal table, λ is the central wavelength, f is the acquisition frequency and n is the refractive index. This means that f will also dictate the measurable velocity range. The settings when acquiring data have to be adjusted so that the velocity of the system falls in the right range otherwise no signal will be detected [11, 21, 35, 36, 39].

Optical Coherence Tomography setup

The OCT-apparatus used in this study is a Telesto II from Thorlabs, which is a spectral domain OCT (SDOCT) with a central wavelength of 1310 nm and a bandwidth of 270 nm. The resolution of the images is less than 3 × 3 × 3µm per voxel. The acquisition frequency for 3D-imaging is 5.5 kHz and 28.8 kHz for 2D Doppler. Because of the angle of the channel the acquisitions for 3D (structural) data are made in 1 mm sections to make sure that the image quality (focus) is sufficient in both ends of the acquisition. The Doppler acquisitions1 are made at

1These generate both a phase field and a 2D structural image of the cross-section. The structural data have been used to compare and confirm the measurements of the height and

(19)

Figure 2.3: Doppler OCT measures the axial velocity component v||. The OCT beam is illustrated in yellow [21].

intervals of 0.5 mm and at each point 100 acquisitions are made averaged2. A manual, linear translation stage with an end-mounted micrometer is used to move the channel.

2.3 Post-processing

The raw data from the OCT is a voxel grid with dB-values of the backscattered light, as well as the resolution. The Matlab-code to process the structural data is based on the one used in Lefranc’s internship report on measurements of flow- focusing with OCT [20], and has been adapted to account for the differences, mainly in geometry.

The main steps of the post-processing are to crop the data to remove that which is not the inside of the channel. The two phases, core and sheath flow, are then separated from each other by means of thresholding the intensity. This way the shape of the thread can be displayed as it evolves downstream. The 1 mm-sections are processed one at a time and then assembled into the full length. Because of

width acquired in 3D. However, the acquisition frequency used to acquire velocity data is selected to capture the present velocities, but does not necessarily give the optimum image quality. For the 3D-measurements a lower frequency is used that provides better image quality, but increases the acquisition time.

2Except upstream where the distance is 1 mm

(20)

the angle of the acquisition plane to the actual cross section, (see Figure 2.1d) the measured height, εz/h, of the thread is the height in the acquired cross section, not the plane perpendicular to the channel, see Figure 2.1c. For the 3D-renderings in this thesis, the difference is small enough that the figures will give a good idea of the thread shape in the channel, but for quantitative comparison with simulations the height should be extracted at the same angle as from the experiments.

Finding the shape and length of the area where the core fluid is in touch with the upper wall (”wetted region”) is done separately, and the voxel layer is processed from the top plane down, manually finding the vertical position of the plastic film covering the channel. The phases are then separated through thresholding.

Processing the velocity comes down to two major points. The conversion of phase-shift to velocity by using

v = ∆φ λ f

4nπ sin α (2.2)

as mentioned previously, and to adjust for an incorrect estimation of velocity mag- nitude because of low scattering from the nano-fibril dispersion. While the mea- sured relative velocity variations within each phase are correct, the low scattering requires corrections. The measured velocity fields Umeasured(x, z) are integrated and normalized by the flow rate Q1 of the pumps. This gives the corrected velocity field as:

Ucorrected(x, z) = Umeasured(x, z) Q1 RR

ΣUmeasured(x, z)dxdz (2.3) where Σ represents the thread cross section [10].

All the results are normalized by the width of the channel, h = 1 mm, giving the non-dimensional coordinates y/h, x/h and z/h. The height of the thread inside the channel is denoted εz/h and the width εx/h, see Figure 2.1c.

(21)

Chapter 3

Measurements on cellulose nano-fibril dispersions

3.1 Material

Cellulose nano-fibrils (CNF) constitute the basic building block in trees, plants and some bacteria [22, 26]. They consists of repetitive polymer chains, β-linked d-glucose, forming alternating amorphous and nanocrystalline regions [22]. In the wood cells the CNFs are organized in a lamellar structure with a highly ordered spiraling orientation along the fiber axis, which ultimately gives the tree its good mechanical qualities. This is due to the fact that the fibrils themselves have excellent mechanical properties and when fully aligned in a fiber this can have a specific ultimate strength comparable to glass fibers and stiffness comparable to Kevlar [22, 24, 26]. In other words; a very desirable bio-fiber can be produced.

The extraction of CNFs requires both chemical and mechanical processes. For CNF from wood this usually starts from pulp, where most of the lignin, hemicel- lulose and impurities have already been removed. Different surface properties of the particles, like surface-charge, can also be modified in the processes by using different protocols like TEMPO-oxidization or carboxylation [26].

CNF used in this study

The cellulose nano-fibrils used in this study is TEMPO-oxidized and prepared according to published protocols [14, 31]. The charge density is 660 ± 12µmol g−1 and the dispersion has a concentration of 0.3 wt%. The shear viscosity is presented in Figure 3.1.

(22)

Figure 3.1: Shear viscosity of the CNF used measured with bob and cup in a Kinexus Pro from Malvern.

3.2 Flow-focusing geometries and parameters

Five different channel geometries are used to study the topology and centerline velocity of the thread. Three different metal plates with 1×1 mm channels are used with the inflow angle between the core flow channel and the sheath flow channels (see Figure 3.2) β = 30, 60 and 90, and by reversing the flow direction: 120 and 150 are also tested. The different cases will hereinafter be referred to by their angle β.

The volumetric flow rates of the pumps are set to Q1 = 23.4 ml h−1 and Q2 = 13.5 ml h−1.

For the sheath flow, deionized (DI) water mixed with 10 % milk1 is used. The milk does not change the properties of the sheath flow2 [10] and was introduced because the optical contrast between the CNF and un-seeded DI-water was too low. By seeding the sheath, and not the core, the physio-chemical properties of the dispersion are better preserved.

1Ultra-high-temperature processed, highly homogenized and pasteurized, almost non- perishable milk with 1.5% fat

2The viscosity of the fluid, however, is raised slightly, to 1.17 mPa s but in comparison to the very viscous dispersion the ratio is close to unchanged.

(23)

Figure 3.2: Schematics of the different flow-focusing geometries used in the CNF- experiments.

3.3 Results

The results of the CNF-study are divided in two sections: (i) topology of the wetted region and the thread and the (ii) centerline velocity variations. The wetting topology is taken from the top plane and Figure 3.3a shows the intensity plot for the 30-case. The clear difference in intensity between sheath and core, as seen in the figure, makes for easy thresholding and high reliability. After converting the phase to velocity, and performing the corrections using the pump flow-rates, a velocity intensity field, as shown in Figure 3.3b, is generated. The centerline velocity for each section is extracted as a mean over the black area shown in the middle of the figure.

Topology results

There is a thread formation for all five geometries. The 3D-renderings of them are shown in Figure 3.4. They are similar, but looking closely at the details the different wetting regions are shown in Figure 3.6d. The grey lines show where the sheath-inlet ends for each geometry: 1 h for 90, 1.15 h for 60 and 120, and 2 h for 30 and 150. The height and width of the core plotted over the length of the channel are shown in Figure 3.6. Especially the height for 150 stands out as much lower than the other ones. It has an aspect ratio of almost 1 and the shape is nearly quadratic. The far downstream mean height, width and aspect ratio for respective inflow angle is shown in 3.7. The shapes of the different cross-sections at 10 h are shown in Figure 3.5.

(24)

(a) (b)

Figure 3.3: (a) Intensity image showing the wetted area for the 30-channel. The CNF-dispersion is dark and the sheath flow brighter. (b) Example of the corrected velocity intensity field at y = 10 h downstream for the 60-case. The black square shows the area where the centerline velocity is extracted from. The low velocity areas in the core are the result of poor signal due to low scattering as mentioned in section 2.3.

Figure 3.4: 3D-thread renditions of the thread shape for all five flow-focusing geometries.

(25)

Figure 3.5: The shape of the cross section at y/h = 10 downstream for all flow- focusing geometries.

(a) (b)

(c) (d)

Figure 3.6: The (a) height, (b) width, (c) aspect ratio of the thread (recall Figure 2.1c) as a function of the downstream location, and (d) The shape of the wetting area and the detachment points.

(26)

Figure 3.7: The final width, height and aspect ratios for all channel geometries taken as a mean in the region 10 6 y/h 6 20.

Velocity results

Figure 3.8a presents the centerline velocity for the different geometries from y/h =

−3 to y/h = 20 . Some specifics are also presented separately. The minimum velocity vmin as a function of the angle β, the end velocity3 vend and ∆v/vmin = (vend− vmin)/vmin are presented in Figure 3.8.

3.4 Discussion

Previous studies at KTH have dealt with assembly of CNF into filaments [25] and measuring the order parameter and velocities along the microchannel [5]. More recently the combination of simulations and OCT was used to further increase the understanding of some of the yet unknown dynamics of the system [10, 20].

This parametric study of the effect of channel geometry is a continuation of these efforts, and the data will be used to compare with simulations.

Independent of that some conclusions could be drawn from the results presented here. There is a stable thread forming in the channel for all five geometries. The threads detach from the upper and lower walls at different positions, and with different wetting region shapes preceding that, but they all detach. For all cases the width first changes rapidly, but seem to reach a constant value after y/h > 6 h, see Figure 3.6b. The height however keeps slowly decaying, see Figure 3.6a. It has been indicated that the width and height are governed by different physical mechanisms [10]. Looking at Figure 3.7, the final height of all cases but 150 are very similar, while the width is smallest for the 90-case and becomes larger for

3Taken as the mean velocity between 17 6 y/h 6 20

(27)

(a)

Figure 3.8: Velocity results, vmin, vmax and ∆v/vmin, for the different inflow angles β.

(28)

the angles both larger and smaller.

The similarities and differences exhibited in the wetting plot, see Figure 3.6d, and the width, might be explained by the momentum components of the sheath flow for the different angles. The 90-case has only a core flow normal compo- nent, while the 60- and 120-cases have the same magnitude core flow normal component, and also close to 90, but their core flow parallel components are in opposite directions, and also smaller than the normal component. For the 30- and 150-cases the relationship between the normal and the parallel components are opposite, with the core flow parallel component being the larger one (and of course in opposite direction for the two cases).

Since the velocity variation along the centerline of the thread will impact the alignment of the fibrils, and therefore ultimately the properties of the final fila- ment, it is of interest to see how the velocity varies for the different flow angles.

The velocity plots show different deceleration and acceleration behaviour for the different angles. Not only does it look like the maximum acceleration changes, but also the extent of the acceleration region, thus indicating a change in inner structure for the five cases in both total alignment as well as alignment for differ- ent length fractions. One can also note that both velocity and aspect ratio of the threads are almost symmetrical with respect to the 90-case.

It has been shown that by changing the geometry of the flow-focus section all the different aspects of the thread properties can be manipulated; indicating that had gelling been introduced in the system, the material properties could have been controlled by the flow geometry.

(29)

Chapter 4

Measurements on protein nano-fibril dispersions

4.1 Material

Whey is a byproduct of cheese-making; it is the liquid remaining after milk has been curdled and strained. From this a mixture of proteins can be isolated, to- gether known as whey protein, which consists of of α-lactalbumin, β-lactoglobulin, serum albumin and immunoglobulins [9]. The proteins self-assemble into amyloid- like fibrils, a structure found in connection with diseases like Alzheimer’s and Parkinson’s, but are also investigated as a building block of new biobased mate- rials, and in this context called protein nano-fibrils (PNF) [37, 38]. The protein β-lactoglobulin forms fibrils when, for example, exposed to high concentrations of urea, in the presence of alcohols, at low pH and at high temperature [38].

Depending on the conditions under which the they are formed, PNFs can be long and straight or short and more curled. As it turns out, these PNFs have very different properties, as well as propensity to form filaments. It is thought to be because of crowding effects and the concentration of the suspension [15], however, the role of peptide hydrolysis as a key to the morphology mechanism has also been investigated [38].

Previous work at KTH [3, 15] tested the spinnability of the different types of PNF. There, no case was found where the straight fibrils formed hydrogel fibers strong enough to overcome the surface tension of being removed from the bath solution the microchannel ends in. The curved fibrils, on the other hand, were able to produce hydrogel fibers that were strong enough to overcome the surface tension.

(30)

Figure 4.1: Shear viscosity for the two types of PNF used. The viscosity of the straight PNF was measured with bob and cup in a Kinexus Pro from Malvern and the curly PNF was measured using cone and plate in a DHR-2 from TA Instruments.

PNF used in this study

In this study two different kinds of PNF are used: one with curled and one with straight fibrils. The curled ones have a concentration1 of 93 mg ml−1 and the straight ones have a concentration of 40 mg ml−1. Their respective viscosities are shown in Figure 4.1.

4.2 Flow-focusing geometries and parameters

In the PNF-experiments two different geometries were explored. Both channels have two flow-focus sections, and the distance between these2 are 2.5 h and 5 h respectively, see Figure 4.2a. The experiments are done for both types of PNF- dispersion on both channels.

The volumetric flow rate of the pumps are set to Q1 = 4.1472 ml h−1, Q2 = 4.976 64 ml h−1 and Q3 = 24.8832 ml h−1.

No gelling agent was introduced in the second sheath, instead a HCl-buffer (10 mM) with the same pH as the fibril dispersion was used in both Q1 and Q2.

1These are theoretical values, meaning the concentration during incubation (fibrillation) be- fore dialysis.

2Measured from the center to center of the sheath flow-channels.

(31)

4.3 Results and observations

The results for the PNF-experiments are limited to structural data and observa- tions of the behavior of the thread in the two geometries illustrated in Figure 4.2a.

The two different dispersions display very different behavior in the channel. In Fig- ure 4.2b the cross-sections at y/x = 0.5, 3, 5.5 and 10 downstream are shown for all four cases. The height- and width development over the channel length for the curly PNF-dispersion is shown in Figure 4.4. The straight PNF-dispersion never properly detaches from the upper and lower walls, although the post-processing of the 2D-Doppler data (presented in Figure 4.2b) is not sophisticated enough to always properly capture the faint outline of the core fluid close to the upper and lower walls for the dispersion with straight PNFs. The width for the straight fibrils is presented in Figure 4.5. From Figure 4.2b one can suspect that that the width of the core in the second focus-section is the same irregardless of the distance from the first focus-section, for curly and straight PNF respectively. And also, that far downstream they reach the same width. This is confirmed in the width plots Figure 4.4a and 4.5.

Figure 4.3 shows the top plane from 3D-acquisitions for the dispersion with straight PNFs for the 5h-channel in the section that encompasses the position y/h = 10. There are, however, two more horizontal lines showing on each side of the thread. In the binarized, unfiltered, cross-sections at y/h = 5.5 and y/h = 10, also in Figure 4.3 there are “arcs” visible on each side of the thread. More clearly at y/h = 5.5, but still visible at y/h = 10. These could explain the extra horizontal lines visible in the intensity image.

Although the thread with curly PNFs stays close to the bottom for both channel geometries, the wetting plots in Figure 4.6 show that after the quick detachment from the top in both cases (a) and (b), the thread shows a constant “snail trail”

for the 5h-channel in (c) but some gaps in (d). Instead the thread in the 2.5h- channel jumps up and down, a motion that was also clearly visible in the preview window of the instrument software and was a consistent feature of the flow in this case. This is also observed in the 2D-acquisitions where 100 cross-section captures in one downstream position over a short time-span shows the fluctuations in the thread position in the channel, see Figure 4.7.

(32)

(a)

(b)

Figure 4.2: (a) The two geometries used and the positions y/h in which cross- section data is taken. (b) Cross-sectional shapes of the PNF topology taken from the 2D-Doppler acquisitions.

(33)

Figure 4.3: To the left, the view of the top plane from 3D-acquisitions for the dispersion with straight PNFs in the section that encompasses the position (y/h = 10). To the right, the binarized, unfiltered, cross-sections at y/h = 5.5 (top) and y/h = 10 (bottom). The “arcs” could explain the three horizontal lines visible in the intensity image.

(34)

(a) (b)

Figure 4.4: (a) Width εx/h and (b) height εz/h of the thread over the downstream position y/h for the dispersion with curly PNFs.

Figure 4.5: Width εx/h as a function of downstream position y/h for the dispersion with straight PNFs

(35)

(a) (b)

(c)

(d)

Figure 4.6: The wetting behaviour for the two double flow-focusing channels. (a) Top view for the 5 h-channel (b) top view for the 2.5 h-channel (c) bottom view for the 5 h-channel (d) bottom view for the 2.5 h-channel.

(36)

Figure 4.7: The thread oscillating up and down in the channel at y/h = 6 for the curly PNF-dispersion in the 2.5h-channel. The time between the images are in the order of 0.3 s. All images are the 1 × 1 h cross-section of the channel.

4.4 Discussion

The different behavior between the curly and straight PNFs captured with the OCT confirms observations made about spinnability of and material properties of the filaments. The dispersion with straight PNFs never detach from the upper and lower walls, staying in the tubing regime, and never forms a thread in the channel, see Figure 4.2b. An explanation for this could partly be the significantly lower viscosity, compared to the dispersion with curly PNFs, see Figure 4.1.

The curly PNFs have been successfully assembled into filaments, but the prop- erties differs with the two channel geometries [27]. Here we see that there is some form of thread formation for both cases. However, in the 5h-channel the core fluid never leaves the lower wall. This means a constant shear along the bottom of the thread and no proper engulfment of the gelling agent that is introduced in the second sheath during spinning. In the case of 2.5h-channel the thread stays close to the bottom but does occasionally detach, as can be seen in Figure 4.6, and oscillates up and down as show in Figure 4.7. The 2.5h-channel has shown better spinning performance, which could be explained by the present observations.

(37)

Chapter 5

Artefacts and errors

In order to correctly interpret the presented data it is necessary to be aware of a number of artefacts. In this chapter these artefacts will be described with the hope that the data will be interpreted correctly. The aspects to be discussed are: image analysis, flow symmetry, bubbles, the milk as tracer, index of refraction variations, pumping oscillations and errors in the traversing.

The actual wetting shapes in Figure 3.6d for CNF and Figure 4.6 for PNF, and points of detachment are extracted from the top planes. Since the acquisitions are 1 mm in length, the wetting region is built from 3 to 4 parts. Each part is cropped individually. This means that an error regarding the angle, as well as the layer chosen, could be present. Manually looking through the voxel layers, from the top plane, to find the first one with a clear outline of the wetted region means sometimes discarding one or two where the region is visible, but too much noise makes post-processing impossible. The error then is in how much the area changes for a wall distance in the order of 3 − 6µm. It could be possible to plot the outline of the wetted region, after post-processing, on top of the intensity image of a voxel layer higher up in the structure, to estimate the difference.

To double check that the flow is symmetrical, the wetting region of the bot- tom plane has been extracted in the same way as the top plane for comparison.

The difference ranges from approximately 0.01 mm to one extreme case of approx- imately 0.3 mm. The flow at the bottom generally detaches a little bit before the top. There is also a secondary flow at the end of the focusing region that is more prominent on the bottom than at the top (see Appendix A). These are all indications that the flow might not be perfectly symmetrical.

There are also bubbles that will make some images non-processable, and that gives slightly different issues for the CNF- and PNF-experiments. In the case of CNF, occasional large bubbles will pass through the core making it expand sud- denly and then shrink to almost nothing (as seen in the captured data afterwards).

The bubbles travel so quickly that for the 3D-data only a few frames in one section

(38)

is affected. These frames can be identified and replaced by the image before or after in the post-processing. The thickness of one of these frames is approximately 3µm. For 2D-Doppler there are 100 acquisitions per position, and if bubbles are passing through the channel only a few of these images will be affected. They are removed and the rest of the set is used.

In the case of the PNF-experiments the problem with bubbles was limited to the dispersion with curly PNFs. Downstream of the thread detachment from the top wall the bubbles in the channel rose to the top wall and created optical distortions as they were positioned between the beam and the thread. This occurred mainly in the region right before the second flow-focusing section where Q3 was significantly higher than Q1 + 2Q2.

The purpose of using milk instead of water for the CNF-experiments sheath is that it gives a better contrast but initial experiments, using a different CNF, indicated a slight discrepancy between using water and using milk. It is however difficult to determine if this is because of some interfacial difference, or due to the slightly higher viscosity of milk, or simply the fact that the poor contrast with water itself gives an error for the case without milk. As the intensity difference between water and the CNF used in this study was too poor without milk it was impossible to investigate this aspect in detail. This is something that might need to be addressed later, but it is beyond the scope of this thesis.

The refractive index used is 1.33, the same one as for water. Other articles mention a refractive index for CNF as around 1.56 [19]. This will affect the height measurement and might also need to be investigated further. The effect on the velocity measurements in corrected by the normalization.

The regular wavy behaviour in the width of the thread (and therefore also in the aspect ratio) of the CNF in Figure 3.6b (and Figure 3.6c) could be because of stick-slip in the sheath flow syringes as the rubber in them seemed to react with the milk. When running them for too long there was a swaying in the thread clearly visible in the camera on the OCT, and the resistance variations could be felt when pushing the syringes by hand. But before it reaches that point a smaller stick-slip behaviour could possibly account for the regular waves.

The traverse of the sample, in all acquisitions, is done by hand using a linear translation stage with an end-mounted micrometer. This can clearly lead to errors in the form of either missing data (if the stage is moved too far) or overlapping data (if the stage is moved too short) between acquisition sections.

(39)

Chapter 6

Concluding remarks

The purpose of the CNF-experiments in this work is mainly to provide exper- imental support to simulations that are in progress at the time of writing this thesis. Aside from that, the experiments show that it is possible to control the cross-sectional shape of the thread as well as the velocity development by altering the micro-channel geometry. Although there are already industrial efforts made to scale up the process of continuous spinning of filaments [1, 23] there are still more things to learn about the physics of the system to develop it further.

The protein provides a whole different class of nano-particles, in comparison to the cellulose. Despite being similar in size, and also being long, slender particles, they exhibit fundamentally different properties and macro structures in nature.

Therefore, the great difference in how the PNF-dispersions develop in the micro- channel is intriguing and is a starting point for future work trying to comprehend and control its mechanisms.

(40)

Appendix A

Secondary flow patterns

The secondary flow in the region at the end of the flow-focus section, where the three streams merge, can clearly be seen in the velocity measurements. In the structural (regular OCT) measurements the addition of milk also makes them visible on the intensity images. They are however more prominent on the bottom of the channel. Possible reasons for this are thermal differences, even though they should be minor; that the feeding of the flow still has an impact as it reaches the focus section or that it is an optical phenomenon that because of the channel angle is only visible on one side and a tilt the other way would reveal them on top instead. The significance of the secondary flow in the images of Figure A.1 is clear and could be an interesting topic for future studies.

(41)

(a) (b)

(c) (d)

(e)

Figure A.1: The images show the intensity measurements in a cross section close to the downstream end of the focus-section. The dark oval in the middle is the CNF-dispersion, the white is the sheath flow and in (a) and (b) the channel walls are visible on the sides. The secondary flow is clearly visible as the striped patterns surrounding the bottom half of the dispersion. The channel angle and exact posi- tion of the image in the channel are (a) 30 at y/h = 1.36, (b) 60 at y/h = 0.94, (c) 90 at y/h = 0.75, (d) 120 at y/h = 0.93, (e) 150 at y/h = 1.70

(42)

Bibliography

[1] Extraterrestrial future for the world’s strongest thread.

https://www.ri.se/en/our-stories/extraterrestrial-future-worlds-strongest- thread. Accessed: 08.05.2019.

[2] S. L. Anna, N. Bontoux, and H. A. Stone, Formation of dispersions using “flow focusing” in microchannels, Applied Physics Letters, 82 (2003), pp. 364–366.

[3] K. Ayaka, Assembly of whey protein nanofibrils into macroscopic filaments, Master’s thesis, KTH, Mechanics, 2015.

[4] C. H. Bowen, B. Dai, C. J. Sargent, W. Bai, P. Ladiwala, H. Feng, W. Huang, D. L. Kaplan, J. M. Galazka, and F. Zhang, Recombinant spidroins fully replicate primary mechanical properties of natural spider silk, Biomacromolecules, 19 (2018), pp. 3853–3860.

[5] C. Brouzet, N. Mittal, L. D. S¨oderberg, and F. Lundell, Size- dependent orientational dynamics of Brownian nanorods, ACS Macro Letters, 7 (2018), pp. 1022–1027.

[6] C. P. Brown, A. D. Whaite, J. M. MacLeod, J. Macdonald, and F. Rosei, With great structure comes great functionality: Understanding and emulating spider silk, Journal of Materials Research, 30 (2015), pp. 108–120.

[7] T. Cubaud and T. G. Mason, Capillary threads and viscous droplets in square microchannels, Physics of Fluids, 20 (2008), p. 053302.

[8] T. Cubaud and T. G. Mason, High-viscosity fluid threads in weakly dif- fusive microfluidic systems, New Journal of Physics, 11 (2009), p. 075029.

[9] H. Farrell, R. Jimenez-Flores, G. Bleck, E. Brown, J. Butler, L. Creamer, C. Hicks, C. Hollar, K. Ng-Kwai-Hang, and H. Swais- good, Nomenclature of the proteins of cows’ milk—sixth revision, Journal of Dairy Science, 87 (2004), pp. 1641 – 1674.

(43)

[10] V. K. Gowda, C. Brouzet, T. Lefranc, L. D. S¨oderberg, and F. Lundell, Effective interfacial tension in flow-focusing of colloidal disper- sions: 3D numerical simulations and experiments, Journal of Fluid Mechan- ics, (Accepted).

[11] I. Grulkowski, I. Gorczynska, M. Szkulmowski, D. Szlag, A. Szkulmowska, R. A. Leitgeb, A. Kowalczyk, and M. Wo- jtkowski, Scanning protocols dedicated to smart velocity ranging in spectral OCT, Optics Express, 17 (2009), pp. 23736–23754.

[12] K. M. H˚akansson, F. Lundell, L. Prahl-Wittberg, and L. D.

S¨oderberg, Nanofibril alignment in flow focusing: Measurements and cal- culations, The Journal of Physical Chemistry B, 120 (2016), pp. 6674–6686.

[13] K. M. O. H˚akansson, A. B. Fall, F. Lundell, S. Yu, C. Krywka, S. V. Roth, G. Santoro, M. Kvick, L. P. Wittberg, L. W˚agberg, and L. D. S¨oderberg, Hydrodynamic alignment and assembly of nanofibrils resulting in strong cellulose filaments, Nature Communications, 5 (2014).

[14] A. Isogai, T. Saito, and H. Fukuzumi, TEMPO-oxidized cellulose nanofibers, Nanoscale, 3 (2011), pp. 71–85.

[15] A. Kamada, N. Mittal, L. D. S¨oderberg, T. Ingverud, W. Ohm, S. V. Roth, F. Lundell, and C. Lendel, Flow-assisted assembly of nanostructured protein microfibers, Proceedings of the National Academy of Sciences, 114 (2017), pp. 1232–1237.

[16] J.-H. Kim, B. S. Shim, H. S. Kim, Y.-J. Lee, S.-K. Min, D. Jang, Z. Abas, and J. Kim, Review of nanocellulose for sustainable future materi- als, International Journal of Precision Engineering and Manufacturing-Green Technology, 2 (2015), pp. 197–213.

[17] D. Kiriya, R. Kawano, H. Onoe, and S. Takeuchi, Microfluidic control of the internal morphology in nanofiber-based macroscopic cables, Angewandte Chemie International Edition, 51 (2012), pp. 7942–7947.

[18] M. Kleman and O. D. Laverntovich, Soft matter physics: an introduc- tion, Springer Science & Business Media, 2007.

[19] E. Lasseuguette, D. Roux, and Y. Nishiyama, Rheological properties of microfibrillar suspension of TEMPO-oxidized pulp, Cellulose, 15 (2008), pp. 425–433.

(44)

[20] T. Lefranc, Exploration of flow focusing through Optical Coherence Tomog- raphy, 2018. Internship report, KTH, School of Engineering Sciences.

[21] R. A. Leitgeb, R. M. Werkmeister, C. Blatter, and L. Schmetterer, Doppler Optical Coherence Tomography, Progress in Reti- nal and Eye Research, 41 (2014), pp. 26–43.

[22] S. Ling, D. L. Kaplan, and M. J. Buehler, Nanofibrils in nature and materials engineering, Nature Reviews Materials, 3 (2018), p. 18016.

[23] M. J. Lundahl, V. Klar, R. Ajdary, N. Norberg, M. Ago, A. G.

Cunha, and O. J. Rojas, Absorbent filaments from cellulose nanofibril hydrogels through continuous coaxial wet spinning, ACS applied materials &

interfaces, 10 (2018), pp. 27287–27296.

[24] F. J. Martin-Martinez, K. Jin, D. Lopez Barreiro, and M. J.

Buehler, The rise of hierarchical nanostructured materials from renewable sources: Learning from nature, ACS nano, 12 (2018), pp. 7425–7433.

[25] N. Mittal, F. Ansari, K. Gowda. V, C. Brouzet, P. Chen, P. T.

Larsson, S. V. Roth, F. Lundell, L. W˚agberg, N. A. Kotov, et al., Multiscale control of nanocellulose assembly: Transferring remark- able nanoscale fibril mechanics to macroscale fibers, ACS nano, 12 (2018), pp. 6378–6388.

[26] R. Moon, A. Martini, J. Nairn, J. Simonsen, and J. Youngblood, Cellulose nanomaterials review: Structure, properties and nanocomposites, Chemical Society Reviews, 40 (2011), pp. 3941–3994.

[27] E. Ornithopoulou. Private Communication, 2019.

[28] J. A. Pojman, C. Whitmore, M. L. Turco Liveri, R. Lombardo, J. Marszalek, R. Parker, and B. Zoltowski, Evidence for the existence of an effective interfacial tension between miscible fluids: Isobutyric acid- water and 1-butanol- water in a spinning-drop tensiometer, Langmuir, 22 (2006), pp. 2569–2577.

[29] M. Razavi, 2 - Bio-based nanostructured materials, in Nanobiomaterials, R. Narayan, ed., Woodhead Publishing, 2018, pp. 17 – 39.

[30] S. Rydefalk, Particle measurements using fluctuations in the regular trans- mittance of light through a particle dispersion : concentration and particle size - theory, measurment principles and applications for pulp and paper pro- duction, Trita-IIP, 2009:05, Skolan f¨or industriell teknik och management, Kungliga Tekniska h¨ogskolan, Stockholm, 2009.

(45)

[31] T. Saito, S. Kimura, Y. Nishiyama, and A. Isogai, Cellulose nanofibers prepared by TEMPO-mediated oxidation of native cellulose, Biomacro- molecules, 8 (2007), pp. 2485–2491.

[32] D. Stifter, Beyond biomedicine: a review of alternative applications and de- velopments for Optical Coherence Tomography, Applied Physics B, 88 (2007), pp. 337–357.

[33] Thorlabs, TELESTO II Spectral Domain OCT System Operating manual, Thorlabs.

[34] D. Truzzolillo, S. Mora, C. Dupas, and L. Cipelletti, Nonequilib- rium interfacial tension in simple and complex fluids, Physical Review X, 6 (2016), p. 041057.

[35] B. J. Vakoc, S. H. Yun, J. F. de Boer, G. J. Tearney, and B. E.

Bouma, Phase-resolved optical frequency domain imaging, Optical Express, 13 (2005), pp. 5483–5493.

[36] J. Walther and E. Koch, Flow Measurement by Lateral Resonant Doppler Optical Coherence Tomography in the Spectral Domain, Applied Sciences, 7 (2017), p. 382.

[37] X. Ye, M. S. Hedenqvist, M. Langton, and C. Lendel, On the role of peptide hydrolysis for fibrillation kinetics and amyloid fibril morphology, RSC Advances, 8 (2018), pp. 6915–6924.

[38] X. Ye, K. Junel, M. G¨allstedt, M. Langton, X.-F. Wei, C. Lendel, M. S. Hedenqvist, and M. Gallstedt, Protein/protein nanocomposite based on whey protein nanofibrils in a whey protein matrix, ACS Sustainable Chemistry And Engineering, 6 (2018), pp. 5462–5469.

[39] Y. Zhao, Z. Chen, C. Saxer, S. Xiang, J. F. de Boer, and J. S.

Nelson, Phase-resolved Optical Coherence Tomography and Optical Doppler Tomography for imaging blood flow in human skin with fast scanning speed and high velocity sensitivity, Optics letters, 25 (2000), pp. 114–116.

[40] B. Zoltowski, Y. Chekanov, J. Masere, J. A. Pojman, and V. Volpert, Evidence for the existence of an effective interfacial tension between miscible fluids. 2. dodecyl acrylate- poly (dodecyl acrylate) in a spin- ning drop tensiometer, Langmuir, 23 (2007), pp. 5522–5531.

(46)

References

Related documents

För att uppskatta den totala effekten av reformerna måste dock hänsyn tas till såväl samt- liga priseffekter som sammansättningseffekter, till följd av ökad försäljningsandel

Från den teoretiska modellen vet vi att när det finns två budgivare på marknaden, och marknadsandelen för månadens vara ökar, så leder detta till lägre

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

Generella styrmedel kan ha varit mindre verksamma än man har trott De generella styrmedlen, till skillnad från de specifika styrmedlen, har kommit att användas i större

Parallellmarknader innebär dock inte en drivkraft för en grön omställning Ökad andel direktförsäljning räddar många lokala producenter och kan tyckas utgöra en drivkraft

Närmare 90 procent av de statliga medlen (intäkter och utgifter) för näringslivets klimatomställning går till generella styrmedel, det vill säga styrmedel som påverkar

Den förbättrade tillgängligheten berör framför allt boende i områden med en mycket hög eller hög tillgänglighet till tätorter, men även antalet personer med längre än

På många små orter i gles- och landsbygder, där varken några nya apotek eller försälj- ningsställen för receptfria läkemedel har tillkommit, är nätet av