• No results found

Non-isothermal capillary flow of plastics related to their thermal and rheological properties

N/A
N/A
Protected

Academic year: 2021

Share "Non-isothermal capillary flow of plastics related to their thermal and rheological properties"

Copied!
177
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Non-isothermal capillary flow of plastics related to their

thermal and rheological properties

Citation for published version (APA):

Lavèn, J. (1985). Non-isothermal capillary flow of plastics related to their thermal and rheological properties.

Delft University Press. http://resolver.tudelft.nl/uuid:506f65d9-200f-4a25-b43a-c0e9bce1dbea

Document status and date:

Published: 28/02/1985

Document Version:

Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be

important differences between the submitted version and the official published version of record. People

interested in the research are advised to contact the author for the final version of the publication, or visit the

DOI to the publisher's website.

• The final author version and the galley proof are versions of the publication after peer review.

• The final published version features the final layout of the paper including the volume, issue and page

numbers.

Link to publication

General rights

Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain

• You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement:

www.tue.nl/taverne

Take down policy

If you believe that this document breaches copyright please contact us at:

openaccess@tue.nl

providing details and we will investigate your claim.

(2)

Non-isothermal capillary flow of plastics

related to their

(3)
(4)

Non-isothermal capillary flow of plastics

related to their

thermal and rheological properties

PROEFSCHRIFT

ter verkrijging van de graad van doctor in de technische wetenschappen aan de Technische Hogeschool Delft,

op gezag van de rector magnificus, Prof. Dr. J.M. Dirken,

in het openbaar te verdedigen ten overstaan van het college van de dekanen

op donderdag 28 februari 1985 te 14.00 uur

door

Jozua Laven

scheikundig doctorandus geboren te 's-Gravenhage

(5)

Dit proefschrift is goedgekeurd door de promotor

Prof. Dr. H. Janeschitz-Kriegl

(6)
(7)

Van velen in mijn omgeving, zowel bij de Technische Hogeschool te Delft als bij het Unilever Research Laboratorium te Vlaardingen, heb ik hulp ge-had of steun ondervonden bij het tot stand komen van dit proefschrift. Hen alien zeg ik hartelijk dank.

Een paar zaken mogen echter niet onvermeld blijven. De heren H. C. Nieuwpoort, R. Verhoeven en G. de Vos hebben een groot deel van de vaak moei lij ke meetopstel I ingen gebouwd. Een aanzienl ij k dee I van de thermofy-sische gegevens dank ik aan nauwkeuring meten door de heer A. Suurland.

De heer J. Verbeek dank ik, behalve voor constructie-adviezen, voor de

buitengewone inzet waarmee hij en de heer H. G. Langer de vele tekeningen hebben vervaardigd. lr. J. de Graaf heeft, als afstudeerder, wezenlijk bij-gedragen aan het totstandkomen van het calorimetrisch capillair. Ing. R. van Donselaar ben ik erkentelijk voor de veelvuldige instrumenteel-fysische ondersteuning.

Van de hulp die ik bij Unilever Research heb ondervonden bij het op papier krijgen van mijn result.at.en dank ik in het bijzonder de heer D.M. Bancroft en Dr. S. de Jong voor hun vele taal kundige adviezen. Niet onver-meld mag verder blijven de redactionele hulp van de heer E.T.J. Eikema en de verzorging van het door de vele formules lastige typewerk door de dames

M.L. Smiet-Kuiken en C. Verwol.

Met genoegen denk ik terug aan de samenwerking en nuttige discussies over dit proefschrift die ik gehad heb met o.a. Dr. lr. A.K. van Aken, lr. D.W. de Bruijne, Dr. lr. J. van Dam, Drs. B. Koeman, Dr. lr. F.H. Gortemaker en Dr. lr. K. te Nijenhuis. De onvoorwaardelijke steun van de promoter bij het afronden van de promotie heb i k bijzonder gewaardeerd.

Tenslotte dank ik naast mijn ouders, die door de geboden opvoeding en scholing deze promotie mede mogelijk maakten, mijn vrouw. Josien, jij weet als geen ander hoe moeilij k het soms was. Zonder jouw opoffering zou nie-mand nu dit boekje lezen.

(8)

CONTENTS

1. INTRODUCTION

2. THE HEAT CONDUCTIVITY OF POLYMER MELTS

2. 1 Introduction

2. 2 Selection of experimental method for heat

conduc-3

3

tivity measurements 3

2.3 Survey of literature on heat conductivity in polymers 6

2.4 Theory of the quasi-steady-state hot-wire method 8

2. 5 Limitations and corrections of the selected method 9

2.5.1 Correction for finite thickness of wire

2.5.2 Correction for the finite thickness of the

9

surrounding medium 10

2.5.3 Correction for the finite length of the wire 11

2.5.4 The onset of natural convection 11

2.5.5 The significance of transfer of radiant heat 14

2. 6 Instrumental

2. 7 Experimental

2. 8 Results and Discussion

2. 9 References

A2.1 Appendix: The Nusselt number of a cylinder in the

centre of a slit

3. THE THERMAL DIFFUSIVITY OF POLYMER MELTS

3.1 Introduction

3. 2 Literature survey

3.3 Theory of the temperature oscillation method 3.4 Instrumental

3. 5 Experimental

3.6 Results and discussion

3. 7 References 21 26 28 32 36 39 39 39 42 45 47 48 52

(9)

4. THE DENSITY AND SPECIFIC HEAT OF POLYMER MELTS; CONSISTENCY OF THE MEASURED THERMOPHYSICAL DATA

4. 1 Introduction

4. 2 Specific heat

4. 2. 1 Methodology

4. 2. 2 Results and discussion

4.3 Density

4.3.1 Methodology

4.3.2 Results and discussion

4.4 Interrelation of the investigated thermophysical

properties

4. 5 References

5. NON-ISOTHERMAL CAPILLARY FLOW: A THEORETICAL

APPROACH

5. 1 Introduction

5.2 Discussion of relevant results from literature

5.2.1 Basic results

5.2.2 Theoretical results

5.2.3 Numerical results

5.3 Velocity and temperature distributions for capillary

flow of a liquid with a temperature and pressure-independent power-law viscosity

5.4 Velocity and temperature distributions for capillary

flow of a liquid with a temperature and pressure-dependent power-law viscosity

5.5 Results: calculations of characteristics of the

de-veloped model

5.6 Discussion of the numerical results of § 5.5

5.6.1 Range of validity of the model 5.6.2 The calculated temperature profiles 5.6.3 Pressure gradient and radial heat flux

55 55 55 55 56 58 58 59 61 62 63 63 64 64 67 70 72 79 83 91 94 95

through the wall 96

5.6.4 The influence of the parameters

a

and~ on

the throughput 97

5.6.5 Testing the model 99

(10)

A5.1 Appendix: Calculation of the first-order corrections on pressure profile and throughput

A5.2 Appendix: Calculation of the temperature and the radial heat flux through the wall, corrected approx-imately to first order

6. THE CALORIMETRIC CAPILLARY: AN EXPERIMENTAL

STUDY OF VISCOUS HEATING EFFECTS

6. 1 Introduction

6.2 The temperature dependence of the viscosity 6. 2. 1 Time-temperature equivalence principle

6. 2. 2 Experimental determination of the temperature

dependence of the zero-shear viscosity 6.2.3 Literature data on the temperature

depen-dence of the zero-shear viscosity

6.2.4 The temperature dependence of the viscosity at constant shear rate

6.3 The pressure dependence of the viscosity

6.3.1 The influence of pseudoplasticity

6.3.2 Experimental methods

6.4

6. 3. 3 The value of the pressure dependence of the viscosity

Flow behaviour of polymer melts under constant shear rate condition

6.4.1 The capillary rheometer

6.4.2 Flow curves: results and discussion 6.4.3 The occurrence of melt fracture

6.5 The calorimetric measurement of energy dissipation in polymer melt flow through a capillary

6.5.1 Earlier attempts to measure viscous heating

by the calorimetric principle 6.5.2 The calorimetric capillary 6.5.3 Calibration and test procedure 6.5.4 Results and discussion

6. 6 References Summary Samenvatting List of symbols Subject index 102 106 109 109 109 109 111 114 116 118 118 118 121 122 122 124 130 132 132 133 136 140 149 151 154 157 163

(11)
(12)

1. INTRODUCTION

The flow of a liquid through a pipe is a classical subject in the science of transport phenomena. This science deals with the transportation of mass and heat. The migration of both of these can proceed by convection as well as by diffusion (or, as it is often called in case of heat, by conduction). Additionally, heat can also be transported by radiation.

In many cases, observations in the area of transport phenomena can be understood with the aid of relatively simple models which, nevertheless, incorporate the essential physical mechanisms. The classical result for the flow of liquids is the one obtained by Hagen and Poiseuille around 1840 for liquid flow through a pipe, which relates the pressure drop to the through-put. Exactly one century ago Graetz published his famous study about heat transfer in pipe flow from wall to liquid and vice versa. This technologically extremely important process has since been the subject of numerous studies in which Graetz' solution was extended, in order to account also for viscous dissipation and for non-Newtonian flow. A historical survey is given in Chapter 5. Most of the studies were performed after 1950. An important motivation for these studies was the development of highly viscous, non-Newtonian, thermoplastic materials. Since computer facilities came available, these heat transfer studies were also extended to industrially important complex geometries like those of extruders and injection moulding equipment. In such systems accurate knowledge of local temperatures is particularly important if thermosetting materials or thermally unstable products must be processed. However, almost no studies are available in which theoretical predictions are directly related to experimental observations. This leads to the situation that people often ignore specific effects in their calculations because they are not really aware of the importance of these effects in practical situations (e.g. the compression heating effect, see Chapter 5). This study is intended to contribute to filling this gap.

(13)

That intention puts several requirements on the present study. Three of them are mentioned here explicitly. First, the experiments should cover industrially relevant processing conditions such as the use of highly viscous, non-Newtonian liquids, flowing at high shear rates under high pressure conditions, while the temperature field is not fully developed. As a model we chose the flow through a circular cylinder. Second, the theoretical model to be tested should be powerful enough to cope with these rather extreme experimental conditions. In principle, two theoretical approaches are possible once the mathematical equations covering all relevant physical aspects are formulated. These approaches are obviously the construction of an analytical solution and the numerical evaluation of the problem. We selected the analyt-ical route because it often gives more insight in how the various physanalyt-ical mechanisms interact and contribute to the experimentally observed behaviour. Nevertheless, we fully admit that a numerical approach also has advantages. For instance, it allows for a rather easy incorporation of the pressure and temperature dependences of physical parameters. In principle, a more or less accurate numerical solution can always be obtained whereas an analytical solution may appear to be unattainable. Finally, our intention also requires accurate knowledge of thermophysical properties like heat conductivity and thermal diffusivity (for instance, an experienced temperature increase is

approximately proportional to the inverse of the heat conductivity).

How-ever, literature values for these quantities show large scatter.

Therefore, this book deals with the following subjects. In Chapter 2 the measurement of heat conductivity in polymer melts is discussed. Results of measurement of this quantity are given as carried out with equipment constructed in our laboratory. Attention is also paid to the increase of heat conductivity due to Boltzmann radiation. In Chapter 3 the thermal diffusivity is dealt with. In Chapter 4 the consistency of our measurements of heat conductivity and thermal diffusivity is checked. Chapter 5 deals with the theoretical analysis of polymer melt flow through a capillary. In Chapter 6, after the rheological characterization of the liquids studied, an equipment by which the viscous heating effect can be quantified is described. Results obtained are discussed in relation to theoretical predictions.

(14)

2. THE HEAT CONDUCTIVITY OF POLYMER MELTS

2.1 Introduction

A quantitative analysis of the heat transfer in liquid flow as discussed in Chapters 5 and 6 requires data on the heat conductivity of the liquid. In as much as data are available for molten polymers, there are rather large differences between data from different sources. Therefore, we present here a study of the heat conductivity of molten plastics. Firstly, we consider the various possible methods together with a review of published investiga-tions. Thereafter, we describe the quasi-steady-state hot-wire and discuss possible sources of error which occur in non-ideal conditions. Finally, we give a description of the experimental set-up together with the results of

our measurements for polystyrene and for low- and high-density

poly-ethylene.

2. 2 Selection of experimental method for heat conductivity measurements The measurement of heat conductivity requires the determination of an energy flux, which is the power that flows per cross-sectional area through a medium under the influence of a temperature gradient. The main problem is in the measurement of the energy flux, not in that of the temperature gradient. The energy flux can, in most cases, only be determined indirectly via the calculation of the electrical power supply. One has to take consider-able precautions to prevent the electrical heat from leaking away along other routes than through the medium under investigation. This difficulty determines, to a large extent, the usefulness of a given geometry (e.g. bar, plate, cylinder, sphere) as a measurement set-up.

From this point of view a system of concentric spheres, in which the inner one is heated, is the best one. Practical problems in connection with alignment, filling of the shell-shaped slit (especially in the case of solid media), energy supply to the inner sphere, seem to form the reasons why this geometry is rarely used.

(15)

After this, the cylindrical geometry has the fewest problems with heat-leakage compensation. As a special case of a cylindrical geometry the "hot-wire" should be mentioned. If this wire is thin enough, axial heat-leakage is neglibible. For this technique a wire of high electric resistance is placed into the medium and heated by an imposed electric current. The temperature rise within the wire or at a distance r of the line source is then measured.

In the case of a flat-plate geometry much heat-leakage can be avoided by placing layers of the medium on both sides of the "hot" plate and en-caging this sandwich between "cold" plates on either side. In this way, the double-platen method is realized. The heat-leakage from the rim of the "hot" plate can, in principle, be ei,iminated with a compensation heater. However the practical realization of this construction is difficult.

Another aspect, which plays a role in the selection of a measuring method, is the shape stability of the medium. With a solid, the flat-plate geometry can be used as well as the bar geometry, in which the temperature gradient is along the bar axis. A concentric cylinder geometry is less suit-able for solids because of the difference between thermal expansion coeffi-cients of medium and measuring eel I. This gives rise to heat contact prob-lems. The influence of a gap developing between medium and cell wall can be minimized by the introduction of a liquid of comparable or higher heat conductivity (with plastics, silicone oil is often used). At very low tempera-tures, at which these liquids also solidify, other solutions have to be found. In those cases one may consider the use of helium gas which has a

reason-ably high thermal conductivity (1). Nevertheless, also at low temperatures

flat plates are preferred to concentric cylinders, provided the plates are flattened well: care has to be taken because plastic solids tend to warp un-der large temperature changes (1). Here, helium gas can be of use as well. With liquids, contrary to solids, a concentric cylinder geometry is more suitable than the plate or bar geometry because the liquid has to be en-closed completely. The walls needed in addition to the hot and cold walls, also conduct heat away from the sample. Especially with plastic melts and organic liquids, the wall conductance is higher than that of the medium it-self. With a suitable choice of the ratio of length to gap width the influence of additional walls can effectively be suppressed with concentric cylinder geometry.

Hence, using the various available geometries, several initial and boundary conditions are encountered. First of all, measurements can be performed in the steady-state, i. e. with temperature and temperature gra-dient being constant with respect to time throughout the medium. With

(16)

quasi-steady-state rneausurements, the temperature gradient is constant but the temperature itself rises continuously with time, either linearly (2) (flat plate) or logarithmically (3) (hot wire). A third possibility is given by the application of what has been called the regular thermal state, often applied by Russian investigators, and described, for example, by Kondratyev and Golubev ( 4). In this method, the inner one of two concentric cylinders (both being initially ;;1t temperature T ) is heated up quickly. Afterwards,

0

the stored heat flows to the outer cylinder, which is thermostated at T 0.

During a part of this cooling process the temperature in each point of the medium obeys the regular-thermal-state condition:

1/(T-T 0) · a(T-T 0)/at = constant.

Generally, a quasi-steady-state measurement (especially with the hot wire) requires less time than a steady-state measurement and thereby has less stringent requirements with respect to temperature control. However, steady-state measurements are more suitable in situations in which the specific heat of the medium changes, which is the case with phase transi-tions.

In a non-isothermal low-viscosity medium, heat transfer is not only effectuated by contacts between adjacent atoms ("pure" or "molecular" condition) but also by natural convection and, with increasing temperatures, also by Boltzmann or black-body radiation. In such cases the hot-wire technique is recommended because

( i) convection is avoided when measurements are taken rapidly after the start of this quasi-steady-state experiment, that is, before the convection is developed, and

(ii) the influence of radiation is minimized when use is made of a very thin wire, the temperature gradient near this wire being so large that the contri-bution of radiation from the wire surface is low due to its small surface area. The interplay of these two phenomena is discussed in more detail in Section 2. 5.

We selected the hot-wire technique for our experimental programme be-cause heat-losses at the wire ends and natural convection are practically absent whereas filling of the measuring equipment does not pose too large problems, notwithstanding the high viscosity of the (melted) samples. The transfer of radiant energy could not be suppressed because the heating wires had to have a minimum thickness for mechanical reasons in view of the highly viscous polymers. However, this is no real drawback, as is explained in Section 2.5.5.

(17)

2. 3 Survey of literature on heat conductivity in polymers

In the past, several investigations have been published concerning heat conduction in plastics, especially since the sixties. This literature will be reviewed here. Generally one has no possibility to judge whether the published data are reliable or not. Consequently, for the majority of publi-cations mention is made only of the measuring method and the materials investigated. Exceptions are made only for a few cases. Discussions on reliability are concentrated in Section 2.8.

In 1958, HolzmUiler and Munx used a steady-state double-platen appa-ratus to measure the heat conduction coefficient A of plastics such as PVC

and PS between 25°C and 80°C (5). In 19591 Cherkasova, applying the

"regular thermal state", measured values of A for PS, LDPE, PU and paraffin wax, in a temperature range from 25°C to 90°C (6). In 1960, Underwood and McTaggart used the quasi-steady-state hot-wire method to measure A of LDPE and PS between 20°C and 200°C (7). Their measuring technique was

relatively primitive and resulted in a rather large scatter of the results (±

10%). In 1961, Kline used a steady-state concentric cylinder apparatus for A-measurements on PS, LDPE, PTFE and on a few epoxy resins between 0°C and 100°C (8). Shoulberg and Shetter determined A of PMMA between 20°C and 160°C with the aid of a double-platen apparatus (9). They noticed large differences in published data on A of PMMA (varying as much as 150%, as based on the lowest value).

In Darmstadt (Federal Republic of Germany), in the sixties, Eiermann,

Hellwege, Hennig, Knappe, and Lohe ~ performed a large number of

apparently accurate A-measurements (on e.g. HOPE, LDPE, PC, PMMA, PS, PTFE, PP, PIB, PVC) at temperatures below and just above the softening temperature by using several versions of double-platen instruments (both steady-state and quasi-steady-state) (10-14), and above the softening temper-ature (between 150°C and 240°C, at pressures between 20 and 300 bar) by using a steady-state concentric cylinder apparatus (15-17). These authors also developed theoretical heat conduction models in order to explain the influence of parameters such as crystallinity, strechting ratio, number and size of side chains, degree of polymerization, degree of cross linking and chain degradation ( 16-23). These models are based on the picture that heat flows from atom to atom and that the type of the bond between the atoms determines the thermal resistance between the atoms. The resistance is small if the spring constant related to such a bond is large. Consequently the resistance of a Van der Waals-bond is approximately ten times larger than

(18)

that of a bond along the main chain of the polymer molecule. A survey of the influence of these parameters in polymer systems has been given in a review by Knappe (24).

Bil and Avtokrativa (25) carried out measurements with a number of thermoplastics and thermohardeners using a quasi-steady-state platen method applying a linearly increasing temperature field, as described by Knappe (2). With the aid of a steady-state double platen apparatus, in 1965, Hansen and Ho carried out a few measurements with PS and PE between 60°C and 160°C (26). A variant of this steady state technique has been developed in

1975 by Hands and Horsfall (33a) who determined f.. of HOPE between ambient

and 250°C. Especially at the highest temperatures, the values given seem to be too low. The physical mechanisms involved and experimental techniques used in thermal conductivity and diffusivity of polymers and rubbers have been reviewed by Hands (33b). Greig and Sahota, in 1979, used a

compara-tive single platen method between - 50°C and + 90°C for the measurement of

PE which had been stretched by extrusion (27). Furthermore, the steady--state concentric cylinder technique have been used in 1965 by Sheldon and Lane (28) (for several types of PE and plastisized PVC between 20°C and 100°C), by Fuller and Fricke (29) in 1971 (for PE, PS, nylon between 150°C and 230°C) and by Ramsey, Fricke and Caskey (30) in 1973 (for well defined commercial PE samples, over the range 160°C-290°C). The results of the latter two investigations show much scatter.

More recently, Cocci and Picot (31) used a hot-wire around which the medium flows in axial direction in order to analyse the influence of molecular orientation on f.. of silicone oil. As outlined before, with axial orientation

one would expect f.. to decrease in radial direction; however, an increase

was found. This may be related to so-called cluster rotation of liquid re-ported by Mooney (32).

In 1977, Karl, Asmussen, Wolf and Ueberreiter measured f.. of PE in an

indirect way, using thermal diffusivity ~'density p and specific heat Cp of

the sample, at T = 178°C and at pressures between 320 bar and 1600 bar

(34). In the rather unconventional method that they used, they suddenly compressed a polymer melt (already being at a high pressure) slightly, resulting in a somewhat higher pressure. The temperature response in the sample, which was kept in a thermostatic vessel, was used to derive a value

of ~; from the maximum temperature rise after compression and from the

measurement of p as a function of temperature at a fixed pressure they calculated C p = T • p-2 · (ap/aT) p · (dT/dP) d' b and thereby a 1a •

(19)

2. 4 Theory of the quasi-steady-state hot-wire method

The quasi-steady-state hot-wire technique involves the measurement of the temperature in or near a line source of heat, embedded in the medium to be investigated. This temperature is monitored from the moment that the line source is heated electrically at a constant rate. In principle the linear line source is considered to be infinitely thin and long and embedded in an infinite medium.

As a first step in the construction of the mathematical solution for the temperature field around a continuous source of heat, the solution for an instaneous line source is presented. Suppose this line source, acting only at t=t

0, is positioned at ( x0, y 0, z), with -oo<z«». Accordingly, the

tempera-ture T in the surrounding medium is described by the differential equation

a

OT

o

2T

o

2T

Clt -

ox

2 +

a/

with the initial condition

T=O at t=t and (x-x )2 + (y-y )2 >

E~

with

e

0

+

0

0 0 0 lJ

and with the boundary condition for the "strength" of the source

+oo +oo

J J

T dxdy = Q at t~t

0

-oo -oo pep

[2.4-1]

[2.4-2a]

[2.4-2b]

where Q is the heat generated in the wire per length. Eq. 2. 4-1 can be

solved via Fourier transformation on x and y. Then, after substituting r2=x2+y2 and taking x

0=y0=0, the classical solution is obtained for the

temperature T at time t and distance r from the instantaneous line source acting at t=t0 with strength Q/(p·Cp):

r2 T(r,t-t 0 )

=

Q e -4a(t-to) [2.4-3] 4ni\.(t-t ) 0

The temperature field around a continuous line source of heat with strength q/(pCP), at time t after the start of heating, follows from

substi-tution of Q by (q ·dt ) in Eq. 2.4-3 and integration over time t between

0 0

(20)

q T =

-4nt..

where the exponential integral is defined as:

co -u e - Ei(-x) =

j

du . u x (2.4-4]

For small values of x, the exponential integral Ei(-x) can be written as Ei(-x)

=

y + ln(x) - x + 1;ix2 + O(x3) where y

=

0,5772 represents the Euler number. In this way for large values of the Fourier number Fo = at/r2 Eq. 2.4-4 becomes:

T(r) =

_q_

j_

y +

1n(

4

a~)+ ~

+ ....

l

4nt..

l

r 4at \

(2.4-5]

By plotting T(r) against log(t), the slope gives 2,3 q/(4rrA.), provided Fo > 25. Whereas Underwood and Mc Taggart measured the temperature at a small distance from the line source in the polymer with the aid of a thermo-couple (7), in the present investigation use is made of a platinum wire for the "hot-wire". This wire could be used both as electrical heating source and as thermometer. In the past this method was applied to low viscosity liquids, for instance by Horrocks and Mclaughlin (35).

2. 5 Limitations and corrections of the selected method

2.5.1 Correction for finite thickness of wire

The result of Eq. 2.4-5 has been derived for a line source, i.e. for an infinitely thin wire, whereas, in practice the wire used has a finite thickness. Carslaw & Jaeger (36, 35) give an expression for the temperature of a cylindrical source with radius R for 4at/R2

»

1:

T

=

~

[-

y + In ( 4at) +

~

+

4rrA. R2 2at

(w-2)R2 (4 t ]

2wat

~-y+ln ;2)~

+ .... (2.5-1]

where w = 2pmcm/(pwCw) is a thermophysical material constant with

sub-scripts m and w denoting the medium and the wire. Eq. 2.5-1 shows that the temperature of a cylindrical source with a finite diameter deviates from that of an ideal line source only shortly after the start of the experiment.

(21)

The influence of w on the temperature of the wire (35) is shown in Fig. 2.5-1. In our experiments such effect is negligible.

2. 5. 2 Correction for the finite thickness of the surrounding medium

The finite radial extension of the medium around the wire causes the temperature field to deviate from Eq. 2.4-5 at long times. For a medium enclosed in a concentric cylinder with radius b, Fischer has derived an equation for the temperature T(r) in the medium (37), which, in view of the previous Section, is only valid for not too small values of time:

[2.5-2]

in which xn are the roots of the zeroth order Bessel function J

0(x), and

N

0(xn) is the corresponding modified Bessel function of the zeroth order.

This equation is only valid if the required number of elements in the series development is limited such that the largest eigenvalue needed ls xn «b/R. For values of at/b2<0, 12 the difference in temperature between the results of Eq. 2.4-5 and 2.5-2 is less than 0,01% but for at/b2=0,25 the difference becomes already 0,4% (35). Such deviation in temperature for a heated wire in a bounded medium is illustrated in Fig. 2.5-1.

4Tt~J -q-2 3

log~

-50 4 5

Fig. 2.5-1 Reduced temperature versus reduced time in a transient hot-wire experiment, in which a wire of radius R is embedded in a medium bounded by an iso-thermal, concentric cylinder surface of radius b. As parameters a series of values of the thermophysical material constant w and of the thickness ratio

(22)

2.5.3 Correction for the finite length of the wire

Horrocks and Mclaughlin have derived formulas for a temperature cor-rection due to the finite length of the line source (35). They discriminate between the abrupt discontinuity in the heating and the heat leakage over the metal wire. On the basis of their theorectical result and our experi-mental observations (Section 2.6) it can be concluded that, with our equip-ment, no correction for finite wire length is needed. Therefore we shall not discuss this problem in more detail.

2. 5. 4 The onset of natural convection

It is stated sometimes that natural convection around thin heating wires may lead to a considerable increase in heat transfer, even with highly vis-cous polymer melts. This problem will be considered here in three different ways. Firstly, it is discussed on the basis of straightforward information, based on correlation functions for natural heat transfer as found in the literature. This will be shown to be irrelevant for our particular geometry,

and thus no conclusion can be drawn. Thereafter, two alternative

approaches (one from literature) will be presented, both indicating that natural convection is negligible in our equipment.

( i) Method based on published correlation functions for natural heat

trans-fer. From measurements of the heat transfer between horizontal concentric

cylinders with diameters D. and D (with D >D.) an apparent heat

con-1 0 0 I

ductivity ·\ can be found which is larger than the real conductivity A. because, apart from conduction, natural convection and radiation (not taken into account here) can also take place. The ratio ~

=

A. /A. > 1 in a steady-a state situation can formally be written (38, 39) as

s

= f(Gr,Pr,D /D.)

0 I [2.5-3]

3 2

where Gr is the Grashoff number Gr = Dig E (Ti-T

0)/V and Pr is the

Prandtl number Pr = v/a = Cp11/A.. Here, g is the intensity of the

gravi-tational field, c, is the cubic thermal expansion coefficient of the medium,

v

and 11 are the kinematic and dynamic viscosities and T. and T are the

I 0

temperatures of the inner and outer cylinders. Neglecting the inertial forces with respect to the viscous forces, Eq. 2.5-3 can be simplified to (39):

(23)

[2.5-4]

where Ra = Gr·Pr is the Rayleigh number.

In estimating

r

for our horizontal-hot-wire apparatus the problem

arises that the existing correlations for ~ (see e.g. Kuehn and Goldstein,

who give a correlation in the form ~ = f(Ra,Pr,D/Di), in which Pr has

only a slight influence ( 40)) almost always have been measured with gases

(Pr~

1). The high level of the Prandtl number for polymers

(~·

108), however, makes the application of these correlations questionable. Another

objection is that the covered range of Ra numbers (from 10-l to 1010) is

not wide enough (40). In the present investigation Ra numbers down to 10-7 are of interest.

The conditions chosen by Kraussolt are more appropriate to the polymer

melt situation because he investigated up to Pr = 3000 (39). However, he

only investigated for D /D. < 3, whereas a value of 200 is representative

0 I

for our equipment. If one ignores the above mentioned objections the way to a conclusion is as follows. According to Kuehn and Goldstein the convective

heat transfer for D /D. = 200 is only slightly different from that for

0 I

D/Di =co (40). For this situation, however, more suitable information is available. Arajs and McLaughlin have given a survey of all published results on natural-convection heat transfer from free horizontal wires ( 41). On the basis of their study we have used the correlation proposed by Morgan (42).

-15 -3 -1 -4

In our experiments Gr ~ 10 (because E ~· 10 K ; Di = 10 m;

t.T

10 K) and

Pr~·

10+8, leading to

Ra~

10-7. According to Morgan,

Nu(Ra=10-7)=0,25 where Nu= hD./.\ = q/{nA(T.-T )}.

I I 0

This convectional Nusselt number should be compared with the one cal-culated for pure, steady-state conduction. As an approximation one may use

the expression for a concentric slit (43): Nu d = 2/ln(D /D.). However,

con o 1

it is of course better to consider the geometry actually used in the present investigation, where a horizontal wire is placed in the middle of a vertical slit with width L, for which we have calculated (see Appendix A2.1): Nucond = 2/{ln(L/D) + 0,247}. Setting D/Di and D/L equal to 200 gives the values 0,377 and 0,364 for Nu con d' which are only slightly higher than

the above mentioned value of Nu for natural convect.ion around a

hori-conv 1115

zontal wire. From the relationship Nut t o a 1 =(Nu · con d+ Nu conv ) , as given

by Kuehn and Goldstein (40), it follows that~= Nutt

1/Nu d ~ 1, which

o a con

would imply that natural convection is negligible. However, we think that in view of all the assumptions and approximations made, no reliable conclusion can be drawn about the importance of convection in our measuring device.

(24)

(ii) Method of Horrocks and Mclaughlin. A better method for estimating the influence of natural convection is given by Horrocks and Mclaughlin (35). These authors start from the conclusion of Kraussold (39) that stationary

free convection in horizontal cylindrical slits is only significant if

Gr0·Pr > 1000, where Gr

0 is defined

o

=

(0

0 -Di)/2. They calculate how far in

as Gr,,

=

o3g £ (T.-T )/V2 with

u I 0

a certain time interval the heat emerging from the line source of their hot-wire experiment has penetrated into the surrounding medium, say over a distance r*. Assuming a quasi-steady-state temperature field within a cylinder with radius r*, they cal-culate at what time after the start of the line source with strength q/pC

3 2 p

the value of Gr*·Pr becomes 1000, where Gr* = o*·E·g (T.-T )/J with

2 6 I 0

o*

=

2(r*-R.) and R.

=

0./2. If at/R.

»

1, the criterion, as developed by

I I I 3 I 1' 2 1'

these authors, can be written as x ·ln(x) = R* with x = (4at)2/(yR.)2 and R* = 2·103 nt..ria/(gqR:Ep ).

t

This criterion, when applied to our sit

1

uation,

I m 5

would al low a measurement time of ·~ 10 s. As the time of measurement in

the present investigation is less than 200 s, one may conclude that, despite some questionable approximations, natural convection is negligible.

(iii) Alternative method. A more elegant method of answering the problem of whether natural convection occurs or not is the following. In a time interval t after the start of the hot-wire experiment (for not too small a time (Sec-tion 2.5.1) and not too long a time (Sec(Sec-tion 2.5.2)) an amount of heat q ·t • L flows out of the wire with length l. As this heat wi II diffuse radially over a certain distance corresponding to an original volume V

0 of the

sur-rounding medium, this volume V will have expanded to V +!:N at time t. At

0 0

time t=O the mass of V was pV . Because the mean temperature rise in V

0 0 0

at time t is qtl/(V pC ) the volume expansion is !:iV = £ qtl/(pC ) . This

0 p I p

represents a loss of mass £qtl/C in the volume V . From this, the driving

p 0

force for natural convection as given by Archimedes' law can never be more

than g£qtl/C , irrespective of whether one considers V or a part of it

p 0

having a local temperature which is higher that the mean one in V

0.

The volume V

0 (or an arbitrary part of it) tends to rise according to

Archimedes. This movement is opposed more or less by two effects, firstly, by the flow resistance around the wire, which for the moment is assumed to be fixed and, secondly, by the shear stress between V

0 and the wall of the

t

According to these formulas, the times after which natural convection

occurs appears to be ~ 1/3 of the times measured by these authors, and

(25)

measuring cell. If one supposes the second effect to be negligible, then the whole upward force is exerted upon the wire. The problem is then exactly equal to that of a horizontal cylinder which, due to gravity, moves

vertical-ly through a liquid. White has calculated the velocity of a horizontal

cylinder, moving vertically in the middle of a vertical slit (44), and ex~

pressed this in a friction coefficient £t which is defined in F = £t·A·pv2/2 where F is the force responsible for the movement and A = 2RL is the

projected area of the cylinder with radius R. He found that

£r= f(Re,L/R,b/R), where 2b is the slit width and Re= 2vRp/r] is the Reynolds number. For small values of Re his result is

with 12,6

c, ·

-log(0,85· b/R) and R R/L

c

2 = 40 - ( 1 + ) . L b/R - 1

According to Jones and Knudsen this result is valid for L/R > 32 and b/R > 9, 17 (45). They checked it for 0,01 < Re < 2. Applying White's result to the present problem, one obtains

2 geqt

v

=

-In the experiments described later in this Section e ~ 10-3 K , -1

q ~ 1 W/m,

c

P ~ 2 • 10 3

JI

(kg· K), ri ~- 104 Pa·s, b/R = 2·102 and

3

L/R~2·10. This implies that, during an experiment of 200 s, the wire will not move more than 8 µm. Taking into account that, in reality, the wire is free to move over such a distance together with the surrounding rubber-Ii ke liquid because it is not stiffly fixed, we must conclude that, also according to this reasoning, natural convection is completely negligible.

2.5.5 The significance of transfer of radiant heat

Heat conductivity is defined as the heat flux under the influence of a gradient in temperature. The temperature is related to the average kinetic energy of the molecules. Temperature differences vanish by kinetic energy transfer between the molecules. According to the kinetic theory of heat conduction this transfer proceeds by particles which transfer an amount of energy (heat) over a given distance (46) (the free path length of such particles). Two types of particles are important in polymer melt systems: molecules and photons (emitted by the molecules). The wave length

(26)

depen-dent energy density of photons in the bulk of a medium is directly given by the Stefan-Boltzmann description of black-body radiation. In liquids, the mean free path length of the molecules (7\m) is of atomic size, say

10-10-10-9 m, while that of the photons (A ) is much larger (Saito and

r

Venant report, for hexane (presumably at room temperature), a value

of N 0, 5 mm ( 47)).

In the remainder of this section, firstly data for lir will be provided, both for polyethylene and for polystyrene. Thereafter, we shall discuss under which conditions and, if so, in what way radiant heat transfer can be treated as part of the total thermal conductivity of a bulk material. Problems arising at boundaries are also discussed. Finally, we shali discuss these effects in relation to the hot-wire cell and make a conclusion about the importance of radiant heat transfer in our measurements.

100 80

HOPE

20 woo 3500 3000 2500 2000 1500 1000 ____.. WAVENUMBEA/!cm"1) 500 100

PS

I 20 I.OOO 3500 3000 2500 2000 1500 1000 500 _ , . . . . WAVENUMBER/!cm-1 )

Fig. 2.5-2 lnfrared transmission spectra of Manolene 6050 (thickness 0,28 mm) and Hostyren N4000 (thickness 0, 16 mm), at 298 K.

200

(27)

·'-F '-re"-e'--"p;;..;a'"-'t-'-h'--'-1 e"""'n"'"'g.,_t'-'h-'-'-'f-'-o-'-r-'-1 R;..;._;_r a;;;..d;;;..i;.;;;at ion . For an estimate of

7i.

r , data on the infra -red transmission were needed. These data were kindly offe-red to us by Mr. F.A. ten Haaft (Koninklijke Shell Laboratorium, Amsterdam) and by Mr. M.A. van Schaick (Unilever Research Laboratorium, Vlaardingen). In Fig. 2.5-2,IR-spectra, taken with a Perkin-Elmer-580 IR-spectrometer, are shown for a high density polyethylene (HOPE) sample (Manolene 6050) and a polystyrene (PS) sample ( Hostyren N4000). These are true transmission spectra, i. e. reflection at the sample surface is suppressed* by placing the polymer sample between KBr windows while using paraffin oil as a contact liquid. Additonally, the sprectra were corrected for the absorption in the window material. In separate experiments, without KBr windows, we investi-gated the influence of temperature on the absorption. It appeared that the absorption changed only a few percent when HOPE and PS were heated from 25°C to 120°c and 105°C respectively. For our purposes we thus assumed that the I R-spectra are independent of temperature.

The mean path length Ar in a non-scattering medium can be calculated with a method described by Rossel and ( 49, 50):

[2.5-1]

where the frequency dependent absorption coefficient av is defined as

av = - log(transmission)/(sample thickness), where

a

is the mean absorption

coefficient and Ev is Planck's result for the emissive energy flux of a black body at frequency v, namely:

2hn2 v3 E\I =

-=---c2 (exp(hv/kT) -1)

The calculated values of 71.r for HOPE and PS at several temperatures are given in Table 2.5-1. In accordance with its higher transmittance, the

* The reflected part of a light beam at right angles to an interface is given by (n1-n2)2 /(n1+n

2)

2 where n

1 and n2 are the refractive indices at both sides of the interface. In our experiments, n(PE) =1,51, N(PS) = 1,59,

(28)

- -4

Table 2.5·1 Values of 1\/(10 m) at several temperatures as calculated with Eq. 2.5-1.

material 298 K 400 K 500 K

polystyrene 2,2 3,8 4,5

polyethylene 16 11 8,3

figures of

7\

for PE are much larger than for PS. The dependence of A on

r r

temperature is due to the shape of the weighting function E11 as plotted

against

v.

The function E.J, which approaches zero at very high and very

low frequencies, has a maximum value at

v

= 2,8215·kT/h thus shifting from

v

= 2,35 · 1013 s- 1 (or w = 784 cm-1) at 400 K to I/= 2,94·1013 s-1 (or

w = 979 cm -l) at 500 K, w being the corresponding wavenumber. The almost

transparent

400 K than

With PS the

character of PE at low frequencies has more influence on Ar at at 500 K, thus resulting in higher 7'i r at lower temperatures. dependence of Ar on temperature, however, is different.

Radiant heat transfer in the bulk as intrinsic part of the thermal conduc-tivity. Whether, for the bulk of a medium, the radiation may be incorporated in the thermal conductivity, depends on the magnitude of fir relative to the size of the system under consideration.

(i) If Ar is relatively small, the so-called "optically-thick" approximation is allowed, provided (dT/dx) is not too large. The radiant and molecular

thermal conductivities 'A and 'A are indiscernable and both contribute to

r

m

the total conductivity: 'At = 'A + 'A . This the case in many practical

situa-r m

tions and also in most devices for the measurement of thermal conductivity. Under such circumstances, 'Ar is given by (46, 51)

[2.5-2]

where a

1 = 5,67 · 10·

8 W·m-2·K-4 is the Stefan-Boltzmann constant. Values

for Ar of polystyrene and polyethylene, using Table 2.5-1, are given in

Table 2. 5-2. At high temperatures, especially, 'A may become a substantial

r

part of 'At. The validity of Eq. 2.5-2 is limited to low values of the tempera-ture gradient. At too high levels of this gradient, i.e. if the condition

(29)

Table 2.5-2 Estimates of the contributions 1-./(W·m-1·K-1) by radiation,to the to.ta! heat conductivity of polystyrene and polyethylene at several temperatures.

material polystyrene polyethylene

l

aT

ox

. Ar.

d.\m

I

.\ m dT 298 K 0,004 0,029

«

1 400 K 0,019 0,049 500 K 0,043 0,072 [2.5-3a]

is not fulfilled, the thermal conductivity cannot be considered constant over

a distance 7i. as is presupposed in the derivation of Eq. 2.5-2. In that

r

derivation it is also assumed that, over a length

A

in the medium, the

r temperature difference !:i.T = T

1-T2 is small enough to justify the substitu-tion of

a

1

(T~-T:)

by 4a1

f

3·ti.T where

T

= (T 1+T 2)/2. This assumption implies a second condition for the validity of Eq. 2. 5-2:

[2.5-3b]

(ii) If 7\r is relatively large the optically-thin approximation is allowed, so the radiation cannot be incorporated in a total heat conductivity as was done in the case of the optically-thick situation. In this case the stationary radiant heat transfer between e.g. opaque parallel boundaries at a mutual distance !:i.x with temperatures T

1 and T 2 and emissivity 0 ~ c. ~ 1 can be approximated with (52)

(2.5~4] In this steady-state case, the importance of radiation is equivalent to an increase of the thermal conductivity by N

ac.T~!:i.x.(1-!:i.x/i\

). This expression

I r

clearly shows that the radiaritnheat transfer cannot be incorporated in the total conductivity because .\r would be dependent on the distance between the plates. Such dependence on !:i.x has previously been demonstrated by

(30)

experiments of Poltz and Jugel (46).

(iii) In the case that the size of the medium and

i\

are comparable, the

situation is more complex. Several attempts have been made to analyse the interaction of radiation and conduction in this case, both in planar (e.g. Refs. 51, 52) and in cylindrical geometry (e.g. Refs. 47, 53). Of these, Refs. 51 and 53 also consider the fact that the absorbtion coefficient (the inverse of the free p.3th length of photons) is a function of the wavelength. One of the reasons for this careful isolation of Am from At as presented in the literature is the verification of theories on the kinetics of molecular motion (see for instance Ref. 54).

Radiant heat transfer at boundaries. Near the boundaries of a medium, the net radiant energy transfer is usually lower than that in the bulk, this re-duction being maximal in the immediate vicinity of a wall, provided that it is not transmitting for light. In case of such an opaque wall, the photons have only one half-space at their disposal. From the kinetic theory of diffusion, one can derive that, if the opaque wall is black, the radiant heat transfer is locally reduced by a factor of two. If that boundary is grey

with emissivity £ < 1, the reduction is more pronounced, the factor being

2/£ (46) (for platinum, from which our hot-wires were made, values of £ between 0, 05 and 0, 2 have been reported ( 47)). This reduction occurs near

the boundaries in a layer of thickness ~ 7i.r. The consequence of this is

that, in practice, the "overall" heat conductivity goes down with decreasing thickness of the slab over which the thermal conductivity is measured ( 46, 51, 55). This effect is negligible, however, if the slab thickness is much larger than 7i.r.

Significance of_ radiation in the hot-wire cell. Because of the advantages

previously mentioned, we selected the hot-wire method for our experiments. This technique can be very useful in minimizing the relative importance of the radiation, by using a wire of only a few micrometer thickness: in the quasi-steady-state situation and at a given temperature of the wire the radiant heat flux at the wire surface is independent of the wire radius, whereas the temperature gradient at the surface, and thus the molecular-conductivity flux, increases with decreasing wire thickness. Hence, this method is well suited for measuring A .

m

For reasons of robustness (see Section 2. 7) we could not use wires

thinner than 10-4 m. Consequently, the radiant energy flux cannot be

(31)

experi-mental set-up, the information given before for the one-dimensional, linear heat flow cannot simply be applied to our case. A thorough numerical analysis of radiation in a hot-wire experiment, using a finite differences method, has been developed by Saito and Vena rt, assuming a frequency independent absorption coefficient ( 47). However, their analysis is re-stricted to

R/i\

< 0,01 where

R

is the wire radius. In our experiments this ratio is 0, 12 for PS and 0,05 for PE. Therefore, their analysis is not appli-cable to our particular geometry.

The way to treat the radiation in our experiments can be deduced on more practical grounds. From the discussion given before on 7i.r' and also from numerical results of Saito and Vena rt ( 47), it can be deduced that, in the course of a hot-wire test, the apparent thermal conductivity (as derived from the slope of the curved line in a temperature versus log (time) plot according to Eq. 2.4-5) increases in a complicated way from a value close to

A at small times to the value A + A at very large times (provided the

m

m

r

boundary of the measuring cell is far enough away). In other words, at times at which the line is curved, the apparent conductivity has a value be-tween A m and A m + A , while at large times, if the line is straight, the r

slope corresponds to A + A . This behaviour parallels the results on A by

m r

Poltz and .Jugel who measured A in a parallel plate geometry under steady-state conditions, as a function of the distance between the plates ( 46).

One would expect the graph of temperature versus log(time) to become straight only when the temperature field has been extended radially over at least a distance 7i.r (or, perhaps better, some miltiple of Ar). Taking at/r2 = 0,5 as a reasonable criterion for the radial extension r of the tem-perature field as a function of time, this leads for r = 7\r to a minimum time of 1,0 s for PS and 3, 7 s for PE (for the thermal diffusivity we used data provided in Chapter 3). In practice, we always found that the lines were curved at small times but became straight at longer times, the transition from one region to the other being at times of 2-3 s in the case of poly-styrene and of 6-12 s in the case of polyethylene. Because (i) we always used the slope of the straight part of the line for the determination of A (the temperature usually was monitored until about 100 s after the start of

a test) and (ii) it can be shown that the requirements formulated in

Eqs. 2.5-3a and 2.5-3b are satisfied in our experiment, we feel confident in claiming that the data to be presented in Section 2.8 are "engineering" values (At)' i.e. they contain both Am and Ar' where the non-negligible

contribution A is defined in Eq. 2.5-2. This is fortunate because in

r

(32)

measured) and in Chapter 6 (where the values of A are used in calculations of the temperature field in capillary flow) we are also interested in 11 engineering" values of A, namely At = Am + Ar.

Because of the frequency dependence of a

11

some precaution is required

with respect to the formulation of conditions for the validity of the 11

optical-ly-thick11 approximation in terms of

A .

However, no theoretical solution

r

would appear to be rivailable which clarifies this point. In view of the fact that curvature in the line of temperature versus log(time) was absent at the large times at which we determined the slope of the line, we assume that the longer free path lengths for the more transparent regions of the

I R spectrum do not invalidate our "optically-thick" approximation.

Z. 6 Instrumental

The measuring cell for the determination of the heat conductivity by the hot wire technique, as shown in Fig. Z.6-1, was built to the specifica-tions of the present author in the laboratory workshop of the Chemistry

Department of Delft University of Technology. It consists of a brass

cylinder A of which the ends are closed by hardened glas windows B

(Silidur, thickness 8 mm). Viton seals, a-rings (C), are used between window and cylinder. To keep the windows in position, screws D press the aluminium rings F against the windows. To avoid excessive stresses (which could arise, for example, from thermal expansion), the aluminium rings F are also fitted with Viton a-rings (G). In the wall of the brass cylinder A,

gas-tight passages for the electrical wire (H

1, Hz) were fitted at four

positions. They worked satisfactory although they did not have a long lifetime.

electrical platinum

Such a passage consists of an aluminium tube containing the

lead-in J fused in a glass bead. Between the passages ~.J, a

wire L was placed as the heating wire with a thickness 10 m.

This was then soldered to the lead-in using special solder (melting point 350°C). This wire (manufactured by pulling a thicker wire through pin-holes) has a very uniform thickness, which is essential for realizing a

uniformly distributed energy dissipation over its whole length

(-~

10-1 m).

Platinum wires (M) of equal thickness were fixed at the passages Hz· The

wires M were electrically spot-welded to wire L, using copper electrodes of electrolytical purity (when copper electrodes of technical quality were used, often the platinum sticked to the electrodes during welding. Thereafter, the

wire L was pulled between the lead-ins H

(33)

(a):

current

(b):

Fig. 2.6-1 The cell for measuring the thermal conductivity with the hot-wire tech-nique. (a): vertical cross-section, (b): horizontal cross-section, (c): passage for electrical wire (symbols: see text).

(34)

it was heated to red-hot for a short period in order to eliminate mechanical-stress-induced microstructural orientations which influence the temperature dependence of its electrical resistance. Furthermore, the measuring cell was equipped with a de-aeration capillary P and a handle (not drawn) by which the cell could be placed in an oil-filled thermostatic bath.

The electrical scheme used is shown in Fig. 2.6-2. The measuring cell A was placed in a Wheatstone bridge, together with two precision resistors B1 and B2 of 5000 ll, one precision resistor C (R = 1 lli Imax = 0, 7 A, tolerance 10 ppm) and a slide potentiometer D (0-8 ll; Imax = 16 A) made of manganin resistance wire (which ensures absence of temperature influence on its resistance at room temperature).

The electrical arrangement described was chosen for the following reason. When an electrical current flows through the measuring cell the platinum wire temperature increases as does its resistance. With a constant voltage over the cell the electrical dissipation would decrease. To avoid this we choose the present Wheatstone bridge design, keeping the voltage over the bridge constant. Resistor D is adjusted so that the sum of the resist-ances C and D equals that of A. When the resistance of A then changes slightly, say by a relative increase

e,

the dissipated power in A is lowered

only by

e

2/4. To realize a constant voltage over the bridge, a DC supply

(35)

G

2 was used which had been equipped with external voltage sensors which

were connected to points

I

1 and

1

2, where the current is supplied to the bridge. With these precautions the power dissipation, which was at a level

of ~ 1 W/m, never changed more than 0, 1% during an experiment.

The voltage control device will, before a test is started, be completely out of its control range due to the fact that at the points

1

1 and

r

2 no

cor-rect voltage is measured. With the power supply G

2 used (Solartron, type

A5-1412), it took 5 s after the start of a measurement for the power supply to become completely stable. To reduce this complication, the voltage sensor leads were connected in such a way, via the magnetic switch P, that before

the start of a test, the voltage was sensored before the mercury switch K.

We used all three parallel connections of switch P to minimize possible contact resistances in the sensor circuit. With these precautions the stabili-zation time of the circuitry is about one second.

The measurement of the platinum wire resistance during a test involves the measurement of both the current and the voltage over the measuring section of the heating wire. These two quantities were measured in two separate tests. The measurement mode (current or voltage) is selected by

switch H

1. The voltage was then determined via the current-free platinum

wires M (Fig. 2. 6-1), the current being measured by a voltage measurement

over resistor C (Fig. 2.6-2). An experiment was started by pressing switch H

2, which closed (i) the magnetic mercury switch K, so allowing the current

to flow through the bridge and which started (ii) the integrator L to

inte-grate a constant voltage signal of DC-supply M

1. This integrated signal is

amplified logarithmically in N and fed, together with either the voltage or

current signal of the cell, into an X-Y recorder giving a plot of the voltage or the current against the logarithm of the time.

With this measuring device the thermal conductivity was determined ac-cording to 4rr ( 1 d R) ( 1 d R )- l q. ;- dT . ;- dlogt 0 0 [2.6-1]

2,303

' A

-where R is the resistance of the wire over the length L. The measurement

of R

0 and of dR/dT was carried out in the following way together with a

series of tests. After rewiring the cell and filling it with polymer, tests were carried out at a series of set temperatures. Before each test the resistance at the set temperature was measured using a 0, 1 V DC calibration source G

(36)

during calibration. The temperature of the thermostatic bath was measured with a calibrated Pt-thermometer in the oil bath. It appeared that R/T is

extremely constant over the temperature range investigated ( ~ 380-500 K),

which implies that, with high accuracy, (1/R

0)(dR/dT) = (1/T).

In order to be sure of proper functioning of the equipment the fol-lowing points were checked, using water-free glycerol as a test liquid: (i) There is no influence of the length of the test section of the heating wire (3 cm vs. 6 cm) on the value obtained for A..

(ii) The value of A. as obtained according to Eq. 2.6-1 is not influenced by the voltage sensing wires M (Fig. 2.6-1) which, in principle, may disturb the cylindrical temperature field around the heating wire. Even welding 5 extra platinum wires (each with length 2 cm) on the test section of the heating wire (~ 6 cm length) did not influence the results.

(iii) The absolute level of the obtained values for A. has been checked by comparing our results in Table 2.6-1 with results of a comparative study of 24 different investigations of A.glycerol' as carried out by Touloukian, Liley and Saxena (55). They calculated a correlation formula on the basis of the eight most reliable sets of data. The mean deviation in their correlation is

1, 7% and the maximum 4,

7%.

In view of the agreement of our results with

their findings we conclude that the absolute level of our A. measurements is very satisfactory.

(iv) In Table 2.6-1 is indicated the time at which a measured plot of voltage or current versus log(time) starts to show curvature due to the onset of convection. Also given is the prediction of this time according to the method of Horrocks and Mclaughlin (35) as described in Section 2.5.4. Experimental times are a factor 2-4 shorter than the predicted ones, this factor being

Table 2.6-1 Values of the thermal conductivity >.. of glycerol (measured and from liter-ature) and of the time tconv after which convection occurs in a hot-wire experiment (both experimental and predicted according to the method of Horrocks and Mclaughlin).

T/°C A.glyc_/(W/m/K) exp, A.gl~c. /(W/m/K) lit t conv. exp. /s tconv. s pred.

1

140 0,303 4,0 9

100 0,303 0,298 8 28

Referenties

GERELATEERDE DOCUMENTEN

Voor de komende vier jaar ligt het accent op verdere optimalisatie van de bedrijfsvoering, identificeren en toepassen van aanvullende high- techelementen en optimalisatie van

Notorhynchus primigenius (Agassiz, 1844) moet zijn: Notorhynchus primigenius (Agassiz, 1835).. De hierboven vermelde rectificaties zijn in

dit bot ooit heeft toebehoord en waar de positie in het lichaam geweest zou kunnen zijn.

4 Hierdie studie poog om aan te toon hoedat sodanige faktore in die uitbouing van munisipale bestuur en in die voorstedelike ontwikkeling van Vereeniging

Voor het monitoren van zuurgraad in habitatgebieden zou de volgende procedure gebruikt kunnen worden: - vaststellen welke habitattypen in principe gevoelig zijn voor bodemverzuring

Figure 6 A view of the Wageningen test garden (left) and the output of the semantic segmentation algorithm (right).. Figure 7 A view of the TrimBot2020 test garden at Wa-

between the two polarizations due to the waveguide birefringence. This work is supported by the Dutch Smartmix-Memphis and Dutch IOP Photonic Devices ... a) Schematic layout of

The research question: “How does the reputation of top executives influence the effect of pro-social CSR initiatives on job selection of prospective job applicants?”