Sei sulla pagina 1di 123

UNIVERSITY OF CALGARY

Rheometric Properties of Pure Liquid Elemental Sulfur

by

Gabriel Oluwaseyi Sofekun

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CHEMISTRY

CALGARY, ALBERTA

AUGUST, 2017

© Gabriel Oluwaseyi Sofekun 2017


Abstract

Much of the transportation and handling of molten sulfur requires liquid sulfur pumps which are

challenging to operate due to the anomalous behaviour of sulfur’s viscosity around its λ-transition

region. Sulfur’s viscosity increases from ≈ 10 cP upon melting at T = 115 °C to ≥ 90 000 cP at

T = 187 °C. While the temperatures for this viscosity transition is well known, no shear related

information can be found in the literature. Understanding the flow behavior of molten sulfur will

allow for better pump design and/or qualification of their operational limits.

The investigation of the viscosity of molten sulfur was performed using a new Anton Paar

MCR 302 rheometer supplied with a 1000 bar titanium pressure cell. Viscosity profiles including

attenuation during heating or cooling are discussed, as well as the rheological behaviour of molten

sulfur encountered at shear rates experienced within pipe flow and/or commercial sulfur pumps.

ii
Acknowledgements

My sincere gratitude goes to my supervisor, Dr. Robert Marriott for his guidance throughout this

project. His passion, insights and knowledge were crucial to the progress of this study especially

when there was a need to modify the rheometer for this study to continue.

Thank you to Prof. Venkataraman Thangadurai and Dr. Simon Trudel for serving on my committee

and their insightful questions during the committee meetings.

My appreciation also goes to the entire Department of Chemistry office and the graduate program

support staff especially Janice Crawford for her administrative guidance as I progressed through

my Master’s degree program. I would like to appreciate the efforts of Edward Cairns, the Senior

Electronics Technician at the Science workshop for his unrelenting help at fixing parts of the

rheometer whenever it developed electronics faults.

My unreserved appreciation goes to Connor Deering for his help and readiness to assist each time

I came calling throughout this study. I also appreciate the assistance of Dr. Fadi Alkhateeb and the

entire members of Dr. Marriott’s Research group for their constructive criticism each time the need

arose. Kyle Wynnyk was very supportive during various stages of putting the experiment set-up

together. Kai Feng, an honours’ research project student from the Qingdao University of Science

and Technology, China, also contributed and worked on this project for six months towards his

Bachelor’s degree.

iii
I would like to recognize the entire staff of Alberta Sulfur Research Ltd. (ASRL) for the varying

nature of assistance they rendered and creating an enabling atmosphere for me during this study. I

am particularly grateful to Francis Bernard, Kevin Lesage, and Patricia Allegri for unique help and

support they offered me throughout my studies.

I am grateful for financial support from Dr. Marriott`s NSERC - ASRL Industrial Research Chair

in Applied Sulfur Chemistry University of Calgary, Department of Chemistry Graduate

Student Award (2015, 2016 and 2017), Faculty of Graduate Studies and the University of Calgary.

To my amazing and darling wife, Oluwakemi Sofekun, I say a big thank you for your love and

understanding through it all. I also say a big thank you to my mum and siblings: Sade, Kole,

Gbenga and IBK for all their support.

iv
To the loving memory of

Uncle Nat Sofekun

v
Table of Contents

Abstract ............................................................................................................................... ii
Acknowledgements ............................................................................................................ iii
Table of Contents ............................................................................................................... vi
List of Tables ................................................................................................................... viii
List of Figures and Illustrations ......................................................................................... ix
List of Symbols, Abbreviations and Nomenclature ........................................................... xi

CHAPTER ONE: INTRODUCTION ..................................................................................1


1.1 Overview of the Thesis ..............................................................................................1
1.2 Physical Properties of Elemental Sulfur ....................................................................2
1.3 Viscosity of Sulfur .....................................................................................................3
1.4 Elemental Sulfur Production ......................................................................................6
1.4.1 Native Sulfur .....................................................................................................6
1.4.2 Frasch Sulfur .....................................................................................................7
1.4.3 Sulfur Recovery .................................................................................................7
1.5 Handling, Transportation and Storage of Elemental Sulfur ......................................8
1.5.1 Safety Concerns in Transporting Liquid Elemental Sulfur ...............................9
1.6 Motivation ................................................................................................................11
1.7 Review of Rheometry ..............................................................................................14
1.7.1 Definitions .......................................................................................................14
1.7.2 Simple Shear Model ........................................................................................15
1.7.3 Flow Behaviours ..............................................................................................17
1.7.3.1 Newtonian Fluids ...................................................................................17
1.7.3.2 Non-Newtonian Fluids ...........................................................................18
1.7.3.3 Shear Thinning (Pseudo-plastic) Fluids .................................................18
1.7.3.4 Shear thickening (Dilatant) Fluids .........................................................19
1.7.3.5 Bingham (Plastic) Fluids .......................................................................19
1.7.3.6 Thixotropic Fluids..................................................................................19
1.7.3.7 Rheopectic Fluids ..................................................................................20
1.7.3.8 Viscoelastic Fluids .................................................................................20
1.7.4 Shear Rheometers ............................................................................................20
1.7.4.1 Capillary Rheometers ............................................................................21
1.7.4.2 Falling Body Rheometers ......................................................................22
1.7.4.3 Drag Flow Rheometers ..........................................................................23
1.8 Choice of Rheometer Geometry ..............................................................................27
1.9 Rheology Models .....................................................................................................29

CHAPTER TWO: REVIEW OF LITERATURE DATA, EXPERIMENTAL TECHNIQUE


AND MODELS ........................................................................................................31
2.1 Sulfur’s Viscosity Measurements ............................................................................31
2.1.1 Measurements Data for the entire Liquid Temperature Range .......................31
2.1.2 Measurements Data below 160 °C ..................................................................35
2.2 Viscosity Models for Sulfur .....................................................................................37
2.2.1 Equilibrium Polymerisation Theory ................................................................37
2.2.2 Modified Reptation Theory .............................................................................41

vi
2.3 Effects of Impurities on Sulfur’s Viscosity .............................................................44
2.3.1 Effects of H2S on Sulfur’s Viscosity ...............................................................44

CHAPTER THREE: EXPERIMENTAL METHOD AND PROCEDURE ......................47


3.1 LabVIEW Code for Pressure Data Acquisition .......................................................47
3.2 Pressure Transducer Calibration ..............................................................................49
3.2.1 Deadweight Tester ...........................................................................................49
3.2.2 Experimental Methods.....................................................................................50
3.2.3 Results and Discussion ....................................................................................52
3.3 Commissioning the MCR 302 Rheometer ...............................................................55
3.3.1 Operation and Hardware of MCR 302 Rheometer ..........................................56
3.3.2 Controls and Data Logging .............................................................................58
3.3.3 Comparing Measured Results with Standards .................................................59
3.3.3.1 Experimental Method ............................................................................60
3.3.3.2 Results and Discussions .........................................................................61
3.4 Experimental Set-up ................................................................................................63
3.5 Materials ..................................................................................................................64
3.6 Experimental Procedure ...........................................................................................65
3.7 FT-IR Analysis ........................................................................................................67

CHAPTER FOUR: RESULTS AND DISUSSION...........................................................68


4.1 Measured Newtonian Viscosity Data of Liquid Elemental Sulfur ..........................68
4.1.1 The FT-IR Spectroscopy Results.....................................................................76
4.1.2 Average Measured Newtonian Viscosity Data of Liquid Elemental Sulfur ...81
4.1.2.1 Re-Optimising the Modified Reptation Model ......................................84
4.2 Non-Newtonian Viscosity Measurements Results of Liquid Elemental Sulfur ......85
4.2.1 Rheology of Sulfur in Mid-Shear Rate Range ................................................85
4.2.1.1 Rheology of Liquid Elemental Sulfur during Pipe Flow .......................85
4.2.1.2 Rheology of Liquid Elemental Sulfur during Pump Operations ...........87

CHAPTER FIVE: CONCLUSIONS AND FUTURE DIRECTIONS ..............................91

...................................................................................................................95

...................................................................................................................98

REFERENCES ................................................................................................................104

vii
List of Tables

Table 1. Advantages and disadvantages of different rheometers. ................................................ 28

Table 2. Relevant parameters for deadweight calibration ............................................................ 53

Table 3. The corrected theoretical pressures, the measured pressures, and uncertainties of the
measured pressures estimated using the standard deviation and a 95.5 % confidence
intervals. ................................................................................................................................ 53

Table 4. Certified values for NIST rheology standard-SRM 2490. ............................................. 60

Table 5. Result of the total analysis of Keyera air prilled sulfur. ................................................ 64

Table 6. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 1st charge of sulfur into the apparatus. ................................................ 68

Table 7. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 2nd charge of sulfur into the apparatus. ................................................ 69

Table 8. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 3rd charge of sulfur into the apparatus. ................................................ 71

Table 9. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 4th charge of sulfur into the apparatus.................................................. 73

Table 10. Total H2S (H2S+H2Sx) content from FTIR analysis of each charge. ........................... 80

Table 11. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures, averaged for the 9th - 14th charges of sulfur into the apparatus. ..................... 82

Table 12. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 5th charge of sulfur into the apparatus.................................................. 98

Table 13. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 6th charge of sulfur into the apparatus.................................................. 99

Table 14. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 7th charge of sulfur into the apparatus................................................ 101

Table 15. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 8th charge of sulfur into the apparatus................................................ 102

viii
List of Figures and Illustrations

Figure 1. (a) S8 - Crown-shape molecules of sulfur consisting eight atoms, (b) • S8 • - diradicals
molecules of sulfur produced as a result of rupturing of the S8 molecules, (c) • Sn • - long
chain diradicals formed from combination of diradicals. ....................................................... 4

Figure 2. Viscosity of liquid sulfur as a function of temperature over the λ-transition region,
empirical model by Shaui and Meisen (1995) of the measurements data of Bacon &
Fanelli (1943). ......................................................................................................................... 5

Figure 3. (a) Deformation of an ideal solid in a simple shear model, (b) Laminar flow of an
ideal liquid in a simple shear model...................................................................................... 15

Figure 4. Common fluids flow behaviours. ................................................................................. 17

Figure 5. (a) Shear stress vs. shear rate profile of common fluid flow behaviours, (b) Viscosity
vs. shear rate curve of common fluid flow behaviours. ........................................................ 18

Figure 6. (a) Concentric cylinder geometry of a rheometer showing its dimensions and the
inner bob; (b) cross-section of a concentric cylinder geometry of a rheometer. .................. 26

Figure 7. Experimental literature data of sulfur’s viscosity. ........................................................ 35

Figure 8. Front panel of the VI to log pressure from the transducer............................................ 48

Figure 9. (a) DWT without weight on the weight stack. (b) DWT with transducer and weight
on stack at equilibrium with pressurized fluid. Notations: ................................................... 52

Figure 10. Deadweight calibration plot of the calculated applied reference pressure (pcorr)
against their corresponding transducer measurements (pmeas). ............................................. 55

Figure 11. Schematic of the experimental set-up for measurements of the rheometric properties
of liquid elemental sulfur. ..................................................................................................... 57

Figure 12. Picture of the experimental set-up for measurements of the rheometric properties
of liquid elemental sulfur. ..................................................................................................... 58

Figure 13. User interface of Rheoplus 4.14 software used for controls and logging rheological
data from MCR 302 to the computer system showing some input variables. ...................... 59

Figure 14. Comparison of viscosity measurements made using the MCR 302 rheometer and
certified values. ..................................................................................................................... 62

Figure 15. Measurements of NIST’s SRM 2490: ........................................................................ 63

Figure 16. Average measured Newtonian viscosity data of molten sulfur as a function of
temperature for the first four charges. ................................................................................... 76

ix
Figure 17. FT-IR spectra of the 1st, 2nd, 3rd and 4th charges showing the total H2S (H2S + H2Sx)
content and peaks. ................................................................................................................. 77

Figure 18. Measurements data of average Newtonian viscosity of molten sulfur as a function
of temperature for different charges...................................................................................... 79

Figure 19. Experimental Newtonian viscosity data and models of molten sulfur as a function
of T ........................................................................................................................................ 84

Figure 20. Measurement data of the η of molten sulfur as a function of γ̇ up to 100 s-1 at
constant T. ............................................................................................................................. 87

Figure 21. Measurement data of the η of molten sulfur as a function of γ̇ up to 1000 s-1 at
constant T. ............................................................................................................................. 88

Figure 22. Newtonian viscosity data and critical rate of shear thickening of molten sulfur as
function of T .......................................................................................................................... 89

Figure 23. Block diagram of the the VI used to log on pressure measurements from the Keller
pressure transducer to a computer system............................................................................. 95

x
List of Symbols, Abbreviations and Nomenclature

Symbol Definition

Acs Specific area


Effective area of the piston at ambient pressure and reference
a0
temperature
Ap Temperature-corrected true effective area

as Entropy (S) term due to end effects

A.U. Arbitrary unit

bd Pressure distortion coefficient

bs Entropy (S) term due to chain lengths

c Coverage factor

CEN European Committee for Standardization

COM Data communication line

CR Shear rate controlled

CS Shear stress controlled

DWT Dead weight tester

d1 Width of suspension at the top of a lever system

d2 Width of suspension at the bottom of a lever system

dh Thickness of a body under shear stress


Distance travelled by the imaginary infinitesimal thin upper layer of a
dl
body under shear stress
Ea Activation energy

F Force

FT-IR Fourier Transform Infra-red

g Acceleration due to gravity

gl Local acceleration due to gravity

G Shear modulus

xi
Ge Plateau modulus of an entangled system

HC Hydrocarbon

HP High pressure
Depth of immersion of an inner cylinder of a concentric cylinder
h
rheometer in a sample
ISO International Organization for Standardisation

k Rate constant

kT Temperature dependent constant

K Apparatus’s constant

Kc Consistency factor

Kx Equilibrium constant for sulfur radical formation reaction

Kp Equilibrium constant for sulfur radical propagation reaction

LabVIEW Laboratory Virtual Instrument Engineering Workbench

l Length of suspension in a pulley system

L Depth of inner cylinder in a concentric cylinder geometry rheometer

Lc Length of capillary viscometer

M Torque

Mav Average molecular weight

MCR Modular Compact Rheometer

mp Mass of a suspended body in a pulley system

Mx Ratio of halogen atoms to sulfur atoms

m Flow index or power law viscosity index

N Polymerisation index

NIST National Institute of Science and Technology

p Pressure

pcorr Corrected deadweight pressure

xii
pmeas Measured pressure from the transducer being calibrated

PRT Platinum resistant thermometer

PID Proportional, integral and differential control action

PV Process variable

Q Volumetric flow rate

𝑟̅ 2 Root-mean-square end-to-end average chain length of a polymer

R Ideal gas constant


Average radius of the inner and outer cylinders of a concentric cylinder
R’
rheometer
Ri Radius of the inner cylinder of a concentric cylinder rheometer

Ro Radius of the outer cylinder of a concentric cylinder rheometer

S Entropy

SP Desired control point

SSE Sum of squares due to error

Sμ Polymeric sulfur

Sπ CS2 soluble sulfur



Sx• Long chain diradical sulfur molecules

T Temperature
Reference temperature at which the properties of the DWT were initially
Tref
measured.
LVE Linear Visco-elastic range

v Velocity

x Chain length of a polymeric chain

yx Mole fraction of sulfur molecules with length x

y‫٭‬ Mole fraction of S8 rings in equilibrium

Zw Weight of average chain length

α Cone angle of inner cylinder in a concentric cylinder rheometer

xiii
αc Coefficients of thermal expansion for the cylinder in a DWT

αp Coefficients of thermal expansion for the piston in a DWT

α-S8 Orthorhombic sulfur

β-S8 Monoclinic sulfur


Height difference between the transducer and the bottom of the piston
Δh
assembly in a DWT
∆H Enthalpy change

∆H4 Enthalpy of propagation reaction

∆S Entropy change

τ Shear stress

τrep Time taken for a polymer chain to exit a tube in a snake-like diffusion
Time taken for a chain of polymerisation index N, to break into two
τbreak
pieces
τf Friction constant

τ0 Yield stress

τw Shear stress at capillary wall

θ Deflection angle

γ Shear strain

γ-S8 Metastable sulfur

γ̇ Shear rate

γ̇ cr Critical rate of shear thickening

γ̇ w Shear rate at the capillary wall


 u Extensional viscosity

η Viscosity

ɳmax Maximum viscosity

ɳM Viscosity of a mixture

ρ Density

xiv
η∞ Viscosity at upper shear limit

η0 Viscosity at lower shear limit

ηp Plastic viscosity

ρair Density of air

ρmass Density of a mass

ρfluid Density of the hydraulic fluid

ρB Density of a falling body in a rheometer

ρl Density of a liquid in a rheometer

ρs Density of sulfur

δ Standard uncertainty

ϕ Polymer weight fraction

ɸv Polymer volume fraction

ω Angular velocity

xv
Chapter One: INTRODUCTION

1.1 Overview of the Thesis

Liquid elemental sulfur (S) undergoes polymerisation at the temperature of its λ-transition

(T ˂ 160 °C). As a result, its viscosity behaves anomalously when compared to typical fluid

behaviour. This means that industrially produced molten sulfur must be pumped within a narrow

temperature (T) range below the onset of its λ-transition at T ≈ 160 °C and above its natural melting

point (T ≈ 120°C).1 Some impurities, such as hydrogen sulfide (H2S), have been observed to

modify the viscosity of sulfur. In this thesis, the viscosity of pure elemental sulfur over its entire

λ-transition region initially studied by R. F Bacon & R. Fanelli in 1943 will be re-examined.2 In

addition, the viscosity vs. shear rate, i.e. the rheology of sulfur, will be reported for the first time.

This data is pertinent to optimal functioning of sulfur pumps and possible injection of liquid sulfur

into depleted underground reservoirs for long term storage. A reasonable future extension of this

work will be to further study the rheology of impure elemental sulfur. In order for these and future

studies, a sensitive high-pressure rotational rheometer (Anton Paar MCR 302) has been modified

and commissioned for this investigation.

In chapter one, the properties of elemental sulfur, the anomalous viscosity behaviour of sulfur

around the λ-transition region and some common processes involved in the production of elemental

sulfur will be discussed. Different rheological techniques and the reasons why a concentric

cylinder rotational rheometric technique was chosen for this investigation are explained. In chapter

two, the available literature on the viscosity of elemental sulfur and the modifying effects of H2S

on the viscosity of sulfur are reviewed. In chapter three, pressure (p) data logging using the

1
National Instruments LabVIEW software, the MCR 302 rheometer with concentric cylinder

measuring geometry, and the experimental set-up and methods for this study are discussed. Finally,

in chapter four, the viscosity and rheological data of sulfur from this study are presented. The

conclusions and the contributions of this research to the existing literature as well as possible future

research directions on this study are provided in chapter five.

1.2 Physical Properties of Elemental Sulfur

Elemental sulfur has the most number of solid allotropes of all elements that have been isolated.3

Over thirty well-defined solid allotropes have been characterized using X-Ray diffraction.4 The

discrepancies in literature regarding the names of sulfur’s many allotropes were partly due to the

numerous crystals that were initially yet to be isolated and identified and the difficulty of purifying

elemental sulfur until Bacon & Fanelli’s sulfur purification method.5,6 Low atomic sulfur

molecules of S2 to S5 are present in gaseous and liquid phases of elemental sulfur but these

molecules have no known crystalline forms. S6 to S15, S18 and S20 homocyclic rings have all been

prepared and isolated in solid form. Larger unstable sulfur rings of S > 25 have also been observed

using various spectroscopic techniques.5

At room temperature, crown-shaped eight-membered rings are the dominant sulfur species with

orthorhombic α-S8 crystals as the stable crystal. At 95.24 °C, the α-S8 crystals transform to

monoclinic β-S8 crystals which are thermodynamically stable relative to the liquid up to

T = 115.18 °C and kinetically stable up to T = 120 °C.1 The monoclinic γ-S8 crystals are metastable

form which have been observed at almost all temperatures below the melt. Just above the

2
thermodynamic melting point, T = 115.18 °C, liquid sulfur contains predominantly S8 ring shape

molecules. As temperature rises these rings begin to rupture to form diradical S8 chains. Note that

sulfur melts naturally at a kinetic melting point near 120 °C, due to the slow disproportionation of

the sulfur rings. At T = 160 °C, the liquid diradical species start to concatenate to form long chain

radicals • S n•, i.e., polymeric sulfur, Sμ. Polymeric sulfur can be isolated from liquid sulfur

equilibrated at temperatures between 120 to 250 °C by rapidly quenching to room temperature and

immediately extracting the non-polymeric sulfur with carbon disulfide (CS2).5

The formation of polymeric sulfur in liquid sulfur at T = 160 °C causes abrupt changes to the

physical properties of liquid sulfur. Properties such as heat capacity, dielectric constant and density

undergo second-order transitions, known as the λ-transition. One of the most pronounced effects

of this phenomenon is observed in the behaviour of the viscosity of liquid sulfur; discussed in more

detail in section 1.3. The colour of elemental sulfur gradually changes from light yellow at melting,

through dark yellow at ca. 160 °C (λ-transition), to a dark brown to red above 250 °C. Sulfur has

low thermal conductivity and vapour pressure throughout its solid and liquid phases. The boiling

point of sulfur is Tvap = 444 °C at atmospheric pressure where the species in the sulfur vapour are

primarily S2, S6 and S8, with S2 being the predominant molecular species.

1.3 Viscosity of Sulfur

Viscosity, ɳ, is a measure of a fluid’s internal friction. At melting (T = 115.18 °C), liquid sulfur is

predominantly ring-shaped S8 molecules and has a viscosity of about 10 - 12 cP. As the

temperature increases, like many other liquids the higher energies of sulfur molecules make them

3
easily glide past one another and therefore, its viscosity decreases with increasing temperature to

a minimum of 7 - 9 cP at 157 °C. Above sulfur’s melting point, sulfur rings begin to rupture to

form diradical chains.7 At 160 °C, increased rupturing of sulfur rings leads to the production of

more of the highly reactive radical species which in turn combine to form chains with variable

lengths (Figure 1).


S8 S8• •
Sn•

T > 160 °C
T ≥ 115 °C
T > 187 °C

(a) (b) (c)

Figure 1. (a) S8 - Crown-shape molecules of sulfur consisting eight atoms, (b) • S8 • - diradicals
molecules of sulfur produced as a result of rupturing of the S8 molecules, (c) • Sn • - long chain
diradicals formed from combination of diradicals.

This polymerisation of sulfur into long chains brings about a sharp increase in the viscosity of

sulfur above 160 °C due to the entanglements of the chains. The effect becomes more pronounced

as temperature increases further, reaching a maximum viscosity of ≈ 93,000 cP at 187 °C, which

is four orders of magnitude larger than the viscosity at melting.2 As the temperature of elemental

sulfur rises above 187 °C, the diradical chains begin to break due to increased thermal

decomposition, leading to reduced chain lengths and viscosity up until boiling point of sulfur at

4
444 °C (Figure 2). As a result of this anomaly, liquid sulfur is usually flowed or pumped within a

safe temperature range of 135 - 155 °C.8

100000
S8 ⇌ •Sx•

80000

60000
ɳ / cP

40000

20000

0
120 140 160 180 200 220 240 260 280 300
T / °C

Figure 2. Viscosity of liquid sulfur as a function of temperature over the λ-transition region,
empirical model by Shaui and Meisen (1995) of the measurements data of Bacon & Fanelli
(1943).2, 9

Certain impurities, such as: hydrogen sulfide (H2S), sulfur dioxide (SO2), halogens, hydrocarbons

and amines, have been found to modify the viscosity of sulfur by limiting the effective chain

lengths.10,11 Many of these impurities reduce the viscosity of liquid sulfur by reactions that shorten

5
the polymer chains of sulfur, where the sulfur diradical chains abstract hydrogen or halogen atoms

in terminal positions thereby shortening the average length of the sulfur chains.12 For example, in

reaction 1-1, H2S reacts with a diradical chain to form polysulfane (H2Sx);

2H2S + ˙S˙x ⇌ HSxH. (1-1)

As a result of the shorter chain lengths, the viscosity of molten sulfur that contains dissolved H2S

is much lower than that of pure liquid sulfur. This also explains why the viscosity of sulfur

contaminated with certain impurities such as hydrocarbons (RH) is lower than the viscosity of pure

liquid sulfur around the λ-transition region, as hydrocarbons can react with sulfur at elevated

temperature to produce H2S and polysulfide (RSxR), as shown in reaction 1-2;

2 RH + Sx → RSxR + H2S, (1-2)

which in turn can decompose to produce CarSul - an insoluble compound of carbon and sulfur.13

1.4 Elemental Sulfur Production

Elemental sulfur can be mined from sulfur deposits close to the ground surface (native sulfur) or

deep-lying sulfur formations by the Frasch process. Currently, nearly all Frasch sulfur has been

replaced by sulfur recovered from the Claus process, which is the partial oxidation of H2S obtained

as by-products from gas purification or hydrodesulphurization of hydrocarbon mixtures in oil

refinery processes.

1.4.1 Native Sulfur

Native elemental sulfur mines can be found trapped in gypsum-anhydrite (CaSO4 ∙ 2H2O - CaSO4)

deposits.14 The mining of this elemental sulfur can be done by open excavation or underground

6
mining depending on the nature of the deposit. Aside from producing elemental sulfur in a pure

form, iron sulfide (FeS) is also mined and subsequently distilled to produce elemental sulfur.

1.4.2 Frasch Sulfur

Frasch sulfur is obtained by pumping superheated water into the matrix of sulfur bearing rocks to

melt the entrapped sulfur. The molten sulfur is then forced to the surface by pumping compressed

air into the deposits. Injection of superheated water is continuous to keep the sulfur in the molten

state. A major challenge in this process is the energy requirement to produce superheated water.

Depending on the complexity of the formation, it can take between 3 - 38 m3 (4 - 50 t) of

superheated water at T = 160 - 165 °C to produce one ton of sulfur, which averages 15.8 GJ energy

demand per ton of sulfur produced.15,16 As a result of the energy intensive nature of the Frasch

process and the abundance of sulfur containing hydrocarbons, operations have since closed and

given way to sulfur recovery from oil and gas production. The increased exploration of sour oil

and gas reservoirs, which were hitherto problematic because of their sulfur content, had

inadvertently led to the replacement of conventional sulfur production through mining by the sulfur

recovered as a by-product of sour crude oil and gas production.

1.4.3 Sulfur Recovery

The original Claus process was a one-step reaction process in which H2S is oxidized with air or

oxygen over a catalyst to form elemental sulfur and water at temperature between 200 - 300 °C.17

2H2S + O2 → S2 + 2H2O. (1-3)

Though highly exothermic, much of the energy given off during this reaction process could not be

harnessed and the sulfur yield was thermodynamically limited to 80 - 90 %.17,18 In order to

7
overcome these short-comings, the modified Claus process was developed and introduced by I.G

Farbenindustrie (ca. 1937).16,17,18 The modified Claus process consists of two reaction steps: the

first step being the highly exothermic partial combustion of H2S to SO2 at about 950 - 1350 °C in

a reaction furnace, from which much of the heat of reaction can be recovered in the form of high

pressure steam from a waste steam exchanger (reaction 1-4). Followed by the catalytic

comproportionation of SO2 and unburned H2S to form elemental sulfur and water (reaction 1-5) at

T = 170 - 370 °C in a catalytic converter:

H2S + 3/2O2 ⇌ SO2 + H2O, (1-4)

2H2S + SO2 ⇌ 3/x Sx + 2 H2O, (1-5)

for an overall reaction:


3H2S + 3/2O2 ⇌ 3/x Sx + 3 H2O. (1-6)

Following the initial reaction furnace, up to four catalytic converters are arranged in series to

ensure yields of sulfur of greater than 99.9 %.17 The elemental sulfur is continually removed from

the condensers after the reaction furnace and each catalytic converter to shift the equilibrium

towards increased elemental sulfur production.

1.5 Handling, Transportation and Storage of Elemental Sulfur

With the production of liquid elemental sulfur, the liquid (rundown) is flowed into holding tanks

or pits, before being transported as liquid, formed into a solid block or other solid products forms

for sales. Blocked sulfur is susceptible to contamination by moisture, soil, rust, or other impurities

during outdoor storage and handling.19 Blocked sulfur can be crushed or re-melted for

transportation as the need arises.


8
When solid sulfur cools, it transitions from β-S8 crystals transforms to α-S8 which are 5% more

dense. This structural destabilization makes sulfur friable.14 The dusts produced when sulfur

crumbles can constitute a safety and environmental hazard as it can easily reach its low explosion

level of 3.5 % volume in air. For example in late 1971, the Vancouver port authority, which is the

port where majority of Canada’s exported sulfur, banned the shipment of crushed sulfur into its

terminal after a major explosion involving sulfur dusts.20

Currently forming elemental sulfur into shapes, sizes and thickness that are more resistant to

attrition are practiced in the sulfur industry. Sulfur prilling, pastillation or pelletizing is a forming

technique used to produce a more satisfactory product for handling and transportation.

About 70 million tons of elemental sulfur (all forms) was consumed globally in 2016.21 Most of

this sulfur was used in the production of sulfuric acid, for which purpose it will be preferred in the

liquid state.22,23 Sulfur is produced in sulfur production facilities in the liquid state and so,

transportation to consumers without solidification can be done, although normally the distance is

too large to viably hold as a liquid. The transportation of liquid elemental sulfur is done using

heated tankers, rail cars, barges, and vessels as well as heated pipelines for short distances.

However, certain challenges are associated with the handling of molten sulfur that makes its large

scale and long distance transportation difficult.

1.5.1 Safety Concerns in Transporting Liquid Elemental Sulfur

There are hazards associated with the equilibrium decomposition of H2Sx in liquid sulfur to H2S

(reaction 1-1). Unlike in most other liquids the solubility of H2S gas in liquid sulfur increases with

9
increasing temperature.12, 24 This is made possible by the reactions of sulfur diradicals present in

liquid sulfur around the λ-transition. As the temperature rises, more sulfur diradicals react with

H2S to form H2Sx. As a result, more H2S gas molecules are absorbed into liquid sulfur.24

When temperature is reduced in molten sulfur facilities (such as holding tanks, rail cars, or

tankers), the reverse reaction (reaction 1-1) is favoured and so, the H2S gas will be slowly released

and saturate the confined headspace or be released upon opening to the surroundings. The 600 ppm

lethal dose (LD50) of H2S can easily be reached should the gas leak out of confined facilities. In

cases where the H2S gas cannot escape directly into the surrounding, there is also a risk of H2S gas

exceeding its lower explosion limit of 3.5 mol % (in air).

Liquid sulfur is effectively transported in pipelines within a narrow temperature range within

which liquid sulfur possesses an optimum viscosity for flow. Most liquid sulfur handling

equipment have a heating mechanism for temperature control and are usually designed to operate

between 135 - 145 °C. In the event that the temperature of sulfur rises above T = 160 °C, the

viscosity of sulfur will increase rendering it impossible to flow. This can lead to pump failure due

to operation at higher load, temperature and shear operation. In converse, if the T of liquid sulfur

inadvertently drops below the melting temperature solid sulfur can cause failure. When this

happens in a pipeline, elemental sulfur will gravitate towards the low bends of the pipeline leaving

behind empty sections. During re-melt operations heat needs to be applied from one end of the

pipeline, otherwise sulfur could melt between adjacent solid sulfur within the pipeline. This would

lead to the expansion of the melt, thereby generating extreme pressures that can rupture the

pipeline.25

10
A number of authors have suggested that elemental sulfur transport using pipeline will reduce the

environmental, safety and operational concerns associated with handling sulfur using other

methods. Estimations made by H. W. Habgood of the Research Council of Alberta suggested liquid

sulfur transportation through pipeline is the most efficient way to transport sulfur compared to

other alternatives based on the increasing production of large tonnage of sulfur and the potential

need to transport elemental sulfur in quantities possibly in excess of 0.5 million tons per year over

distances of 965 km.25,26 Pipelines are still scantily used in transporting liquid sulfur for moderate

to long distances in the industry. Some of the few sulfur pipelines in operation are: the 7 km

pipeline in the north of Germany; the 14.5 km undersea pipeline between Grande Isle, Gulf of

Mexico and USA main land; the 24 km pipeline from the Berri natural gas plant to the Jubail sulfur

terminal in Saudi Arabia; the 33 km pipeline in Carter Creek, Wyoming, U.S.A; and the Shell’s

underground sulfur pipeline spanning a distance of 41 km between Caroline and Shantz, Alberta,

Canada16,25,27,28 Some of these pipelines are electrically heat-traced, while the Shell’s sulfur

pipeline is hot-water heat traced.

1.6 Motivation

The global demand for elemental sulfur has continually been on the rise with the increasing need

to feed the world’s increasing population.15 Sulfuric acid produced from elemental sulfur is used

for the digestion of phosphate in the production of phosphate fertilizers. Using elemental sulfur is

the most viable route to the production of soluble phosphates.15 The increased global energy

demand has led to increased production of sour oil and gas which in turn has increased sulfur

recovery. Environmental and safety concerns in addition to strict government legislations have

11
also contributed to the unprecedented amount of elemental sulfur inventory. For example in North

America, the maximum allowable dissolved H2S in liquid sulfur is 10 ppm and the average Claus

recovered sulfur contains 200 - 300 ppm dissolved H2S.18 Hence, the need to further degas the

produce sulfur of dissolved H2S which may be recycled as feed to a sulfur recovery unit or directly

converted to elemental sulfur.

In order to accommodate the potential future over production of sulfur, suggestions have been

made about the long-term storage of sulfur.23 Acid gas injection is already practiced for the

disposal of sulfur in the form of H2S; however, there are strict regulations around the practice and

at times suitable injection reservoirs may not be available for this purpose. Conversion of H2S to

sulfuric acid is another possible sulfur storage method but the toxicity, corrosive nature and its low

sulfur density makes it more expensive to transport and store than elemental sulfur.

Much of the transportation and handling of liquid sulfur requires sulfur pumps which may be

challenging to operate due to the anomalous behaviour of sulfur’s viscosity around its λ-transition

region. High shear deformation will be experienced in pumps as liquid flow within the pumps. At

high temperature and high shear, sulfur pumps can fail. At present, the viscosity versus shear rate

data, i.e., rheometric data of liquid sulfur, is not available in open literature.29 The availability of

this data will help to optimize the functioning of liquid sulfur pumps and improve the handling

process of molten sulfur in general. One of the goals of this thesis is to better understand sulfur’s

flow behaviour.

12
Despite being a major safety concern, the presence of dissolved H2S in liquid sulfur is an advantage

to the flow properties of liquid sulfur. The reaction of H2S with sulfur diradicals reduces the chain

lengths of the diradicals causing a reduced viscosity of liquid sulfur. The modifying effects of

some impurities such as H2S and halogens on the viscosity of liquid sulfur have been reported in

literature.10,11 Bacon & Fanelli (1946) have measured the modifying effects of H2S on viscosity of

liquid sulfur after treatment with H2S gas but these measurements were not correlated with the

actual concentration of dissolved H2S in liquid sulfur.2 Providing a correlation between dissolved

H2S concentration and sulfur viscosity will be a future goal for this type of research.

In comparison to other forms, liquid elemental sulfur is benign, non-toxic and safer to handle

because of its low vapour pressure throughout its liquid state.30,31 As a result, the injection of liquid

sulfur into depleted reservoir has been suggested as a safe alternative for long term storage of

elemental sulfur. In order to achieve this injection there is again, the need to define the flow

properties of liquid sulfur, i.e., the rheometric properties. In addition to the high shear deformation

experienced within pumps, high shear will also be experienced near the well-bore region of

reservoirs. It is therefore not sufficient to know the viscosity of liquid sulfur without relating it to

shear rate. The polymerisation of sulfur around its λ-transition suggests that liquid sulfur may

possess non-Newtonian properties like most polymers, and so a thorough understanding, the

mitigation of the anomalous behaviour of liquid sulfur’s flow property may be achieved by

investigating sulfur’s viscosity with respect to shear, i.e., its rheological properties.

13
1.7 Review of Rheometry

1.7.1 Definitions

Rheology is the study of flow and deformation of materials under the influence of stress. An ideal

solid, also known as a Hookean solid, obeys Hooke’s law and deforms reversibly in response to

an applied stress.32 The energy expended for its deformation is fully recovered when stress is

removed. Conversely, an ideal liquid obeys Newton’s law and deforms irreversibly. The energy

expended for the deformation of an ideal liquid is dissipated within the liquid in the form of heat

and is not recovered when the stress is removed.33 In other words, the response of an ideal solid to

shear stress is a reversible elastic deformation while that of an ideal liquid is an irreversible viscous

flow. Most materials of practical importance are neither ideal solid nor ideal liquid and are

therefore commonly referred to as viscoelastic materials.33 Their rheological properties fall in

between that of ideal solids and ideal liquids.

At early stages of its development, rheometry was defined as standard techniques for measuring

shear viscosity. However, as interest in non-Newtonian fluids (fluids whose viscosity change with

respect to shear rate) grew, rheometry expanded to include techniques such as measurements of

normal stress and elongation viscosity.34 The modern definition of rheometry includes techniques

in which viscosity and viscoelastic properties of a material are studied as a function of temperature,

time, shear rate, or stress while the material is subjected to controlled temperature, shear rate, or

stress program.35 Whereas, in viscometry, viscosity is measured as a function of temperature, time,

shear rate, or stress while the material is subjected to a constant temperature, shear rate, or stress

program. Viscometry is therefore a branch of rheometry.

14
1.7.2 Simple Shear Model

In order to establish the basic parameters necessary to define the flow properties of materials, a

simple shear two parallel-plate deformation model is used to mimic a simple laminar flow regime

in both solids and liquids.

Shear Strain,
τ = (F/A) Shear Stress,
dl
Uniform Velocity, v τ = (F/A)

dh
h 0

Area (A) dv

(a) (b)

Figure 3. (a) Deformation of an ideal solid in a simple shear model, (b) Laminar flow of an ideal
liquid in a simple shear model.

In Figure 3a, an ideal solid is subjected to a shear stress, τ, (defined as the force, F, per unit area, A)

tangentially applied in the direction of shear (deformation). When F is applied to the solid with

thickness, h, the top of the solid moves a certain distance dl, then stops in the new position for as

long as the force is applied. The response of the solid material is defined by the shear strain,

γ (dl/h). The shear modulus G, (equation 1-7) is a measure of stiffness linked to the chemical

make-up and physical nature of the solid under investigation.32 It defines the resistance of the

solid to deformation:

dl
  G  G  (1-7)
h

15
In a liquid, the shear stress, τ, applied to the liquid acts tangentially to the surface of the liquid as

if it were stacks-of-cards sandwiched between a moving upper and fixed lower parallel plates

(Figure 3b). As a result, a sliding of one infinitesimal layer, dh over another layer is produced in

the liquid.36 When a shear force is applied to the surface of the liquid, the upper plate moves at a

constant velocity, v, for as long as the force is applied. When the shear stress is removed the plate

stops moving and retains its new position (unless the inertia due to its mass continues its motion

for a short time). The velocity decreases linearly across the thickness (h), between the moving

upper plate and the fixed lower plate. Similar to the solid, a shear strain rate or shear rate,

γ̇ (dv/dh), is produced in the liquid as a result of the shear stress applied. In an ideal liquid, the

shear rate is proportional to the shear stress and the constant of proportionality is the viscosity,

η (equation 1-8). This is a measure of the resistance a liquid possesses against any irreversible

positional change of its volume elements.

dv
       (1-8)
dh

Flows in liquids can generally be divided into extensional and shear flows. For extensional flow,

adjacent elements flow away or towards each other, whereas in shear flow, liquid elements flow

past each other.37 There have been little extensional rheological measurements available in

literature due to the difficulty involved in their measurements. However, the Trouton rule

(equation 1-9), had been widely used to predict uniaxial extensional viscosity of a fluid, ηu, from

its viscosity, η values.

u  3 (1-9)

Experimental data later showed that extensional viscosity could be a lot larger than predicted by

Trouton’s rule at higher shear rate.38 Henceforth, no attention will be given to extensional

16
Rheometry in this thesis because in the liquid sulfur processing facilities, such as sulfur pumps

and drainage pipes, shear flow is the dominant flow and this thesis seeks to establish flow

parameters that would address problems in such facilities.

1.7.3 Flow Behaviours

A material’s response to shear stresses can generally be characterised as Newtonian or non-

Newtonian. Non-Newtonian material behaviour is further divided into various groups as shown in

Figure 4:

Figure 4. Common fluids flow behaviours.

1.7.3.1 Newtonian Fluids

Fluids that deform uniformly with respect to shear rate are defined by Newton’s law of viscosity,

equation (1-8). The viscosity of these fluids is independent of shear rate and are referred to as

Newtonian fluids i.e., an ideal fluid. Liquid water is an example of a fluid that exhibits this flow

behaviour.

17
1.7.3.2 Non-Newtonian Fluids

Many fluids that have industrial applications do not shear uniformly and cannot be defined

perfectly by Newton’s law of viscosity (Figure 5). The viscosity of these fluids vary with shear

rate and /or time and are further classified as: shear thinning, shear thickening, Bingham,

thixotropic, rheopectic, and visco-elastic fluids.


Shear stress, τ / Pa

Viscosity, η / cP Newtonian

Bingham

Yield stress, Shear rate, γ˙ / s-1 Shear rate, ˙γ / s-1


τ0 / Pa

(a) (b)

Figure 5. (a) Shear stress vs. shear rate profile of common fluid flow behaviours, (b) Viscosity vs.
shear rate curve of common fluid flow behaviours.

1.7.3.3 Shear Thinning (Pseudo-plastic) Fluids

Shear thinning behaviour is a common flow behaviour in materials. The viscosity of shear thinning

materials decreases with increasing shear rate (Figure 5). At rest, these materials contain randomly

distributed particles/molecules with irregular shapes and sizes. As flow begins, these particles are

forced against one another causing moderate resistance to flow.33 But, as shear rate increases, the

particles become aligned to the direction of flow and as a result, the viscosity reduces until it gets

to a constant value where there is little or no interaction among material particles. For example,

18
squeezing ketchup out of a flexible bottle increases the shear rate and reduces the viscosity of the

fluid making more of the ketchup flow out of the bottle.

1.7.3.4 Shear thickening (Dilatant) Fluids

Some materials undergo an increase in viscosity as the shear rate increases (Figure 5). The

application of shear force to this group of materials causes layers to slide against each in other to

flow. As shear rate increases, irregularities in particle ordering results in increased resistance to

flow and eventually leads to particles jamming together.39 Suspensions of corn starch in water is a

typical example of a shear thickening fluid.

1.7.3.5 Bingham (Plastic) Fluids

Bingham flow behaviour is similar to shear thinning behaviour except that a certain shear stress

value known as yield stress, τ0, must be attained before flow can commence at all for fluids that

exhibit this behaviour, (Figure 5a). At rest the microstructures of these materials contain networks

that greatly resist deformation like a solid. These networks are broken down at the yield stress,

which is unique for the material, before viscosity can start to decrease with increasing shear rate.

Toothpastes and gels show this flow behaviour.

1.7.3.6 Thixotropic Fluids

This is one of the two time-dependent flow behaviours in fluids. Thixotropic materials undergo

shear thinning as a function of shear rate and time. The shear history of the materials influences

their flow behaviour. If the shear rate is varied in a cycle of an increasing shear rate followed by a

decreasing shear rate, the viscosity vs. shear rate curve obtained will be anisotropic, indicative of

19
partial recovery of the molecular structure of the materials. An example of a thixotropic material

is cement paste.

1.7.3.7 Rheopectic Fluids

This is a time-dependent shear thickening behaviour in materials. Rheopectic materials show an

increase in viscosity as a function of shear rate and time. The viscosity vs. shear rate curve is also

anisotropic, though in a counter-clockwise manner, i.e., the downward viscosity vs. shear rate

curve is above the upward curve. Essentially, the structure of rheopectic materials break down at

rest and is built up while being subjected to shearing. For example, 42 % aqueous gypsum paste

at rest solidifies after 20 s if gently rolled out of a container onto a surface.40

1.7.3.8 Viscoelastic Fluids

A viscoelastic materials’ response to shear stress is both elastic and viscous. Discrepancies exist

in literature in designating materials as an ideal solid, ideal liquid or viscoelastic. This is because

the response of a material to shear depends on its structure and the condition it has been subjected

to. A dynamic shear stress experiment is often carried out to establish the strain within which the

response of a material is both elastic and viscous. A material property obtained is known as linear

viscoelastic limit (LVE).41

1.7.4 Shear Rheometers

In order to classify the flow behaviour of a fluid, shear rheometers are often used to measure

rheometric properties. They can be divided into: capillary rheometers, falling body rheometers and

drag-driven flow rheometers.

20
1.7.4.1 Capillary Rheometers

Capillary rheometers are also referred to as pressure driven flow rheometers or Poiseuille

rheometers. It is worthwhile to mention here that for Newtonian fluids, viscosities are normally

obtained in capillary viscometers in terms of kinematic viscosity, ν, (ɳ/ρ). Because the force of

gravity is oftentimes the driving force that pushes the liquid through the capillary, then the density

(ρ) of the liquid becomes an additional parameter and is used to convert measured kinematic

viscosities to dynamic or absolute viscosities. In capillary rheometry, a plunger, an extruder, or

any other source of pressure drives the fluid through the capillary either at a constant flow rate (Q)

or with a predetermined pressure drop (∆p). The flow rate is obtained by timing the movement of

plunger or extruder. In capillary flow rheometers, the Hagen ̶ Poiseuille expression (equation 1-10)

applies, assuming a steady, laminar and isothermal flow.36 where r is the radius of capillary, ∆p is

the pressure drop through the capillary, V is the volume of liquid that flows in time t, and Lc stands

for length of the capillary tube:

r 4 pt
 . (1-10)
8VLc

The shear rate at the capillary wall, γ̇w and shear stress at the capillary wall, τw are obtained from

equations (1-11) and (1-12) respectively, where Q represents the volumetric flow rate;

4Q
w  , and (1-11)
r 3

rp
w  . (1-12)
2L

Equation (1-10) can be reduced to ɳ = Kt, for a particular capillary rheometer. K is the instrument

constant obtained by measuring the viscosities of standard samples using the instrument. The

21
viscosity of any sample under investigation is determined by multiplying K with the efflux time, t.

The pressure losses as a result of the turbulence and elasticity witnessed at the entry and exit of

capillary are referred to as end-effect losses. These are corrected for, depending on the capillary

instrument. In addition, because a velocity gradient exists from the center of the capillary to the

wall, the non-linear velocity gradient is corrected for by multiplying the η in equation (1-10) with

the Rabinowitsch correction factor (3m + 1)/4m to convert the apparent shear rate to a true shear

rate where m is the power law coefficient which is an experimentally determined material

property.36

Capillary Rheometers for Use with Molten Sulfur

Sulfur undergoes polymerisation reaction around T = 160 °C, which makes non-newtonian

behaviour probable, as most polymers have non-Newtonian characteristics.36 The fact that a new

volume element flows through the capillary over some time, makes it impossible to define shear

for an extended time in capillary rheometers. In addition, thermal degradation and equilibration of

polymeric sulfur are time dependent and are at the crux of this thesis and so, capillary rheometers

cannot be used in this study.

1.7.4.2 Falling Body Rheometers

Some examples of falling body rheometers are: falling sphere, falling needle, and sliding ball. For

example the design of falling sphere rheometer is based on the Stoke’s law, which relates the

viscosity of a fluid to the velocity of the sphere falling through it. In equation 1-13, r, is the radius

of the sphere, ρB and ρl are the densities of the falling body and the liquid respectively; g is the

acceleration due to gravity and v is the velocity of the falling body:

22
2 gr 2 (  B  l )
 (1-13)
9v

These instruments are rarely used for rheological measurements due to the difficulties involved in

accurately correcting for non-Newtonian behaviour.38

1.7.4.3 Drag Flow Rheometers

In drag flow rheometers, the motion of a moving surface in the rheometer causes the shearing of

the fluid sandwiched between the moving surface and a stationary surface. The fluid imparts a

drag on the stationary surface of the equipment which is measured alongside the velocity of the

moving part. The shear rate and shear stress are respectively obtained from the velocity of the

moving part and the force per unit area required for the motion. The viscosity and other rheological

data are calculated from these parameters based on the geometry of the drag flow rheometer as

highlighted below.

Sliding Plates Rheometers

A rheometer with parallel plates’ geometry is a direct development from the simple shear model

(Figure 1b). A steady simple shear is generated by placing materials between a fixed plate and

another plate moving at a constant velocity.38 In these instruments, shear stresses are derived from

the force acting to displace the moving plate or keeping the stationary plate in position. These

instruments are equipped with shear stress transducers to measure local shear stresses.42 Though,

these rheometers are more suitable for measurements of viscosity at high deformation rates, the

challenges of operating them include errors due to edge effects and the difficulty of confining less

viscous fluids within sample space.

23
Rotational Rheometers

There are two methods to induce shear in a rotational rheometer. One way is to pre-set a rotational

speed in the instrument and then, the torque generated by the fluid is measured. The other way is

to impact the fluid with a known torque, then the rotational speed response is measured. At the

advent of rotational rheometric techniques, instruments that were designed to operate base on the

pre-set rotational speed were referred to as controlled rate rheometers (CR), while those that were

based on the pre-set torque were called controlled stress rheometer (CS).33 In most cases, as a

result of the power of computing, these instruments have long been replaced by more sophisticated

designs that could operate in both CR and CS modes. In addition, many new rheometers are now

modular, thus users can easily switch from one measuring configuration to another on the same

rheometer.

There are three basic shear induced flow models in laminar flow regimes upon which the designs

of rotational rheometers are based. Below are the descriptions of the three models vis-à-vis their

geometries and instrument designs.

Parallel Disks

In this design, one of the two parallel circular plates with small gap between them rotate around a

central axis as the fluid under investigation is sandwiched between them. Either the upper or lower

plate can serve as the stationary plate while the other will be the moving plate. A laminar flow is

assumed, similar to sliding motion in a pack of cards. The major challenges in the use of this

geometry are the edge effects and fluid inertia. The edge effects which may be as a result of

24
evaporation or aeration of samples at the edges of plates. To minimise edge errors, the gap (h)

value must be at least fifty times the average dispersed particle diameter.32,39 Fluid inertia depends

on the nature of material under investigation. Both of these sources of error are concerns in

measuring the rheometric properties of sulfur and will greatly reduce the accuracy of

measurements. Finally, because the studies within this thesis are aimed at a wide temperature and

longer periods of investigation, fluid loss due to evaporation would be problematic.

Cone and Plate

Because a constant velocity gradient is assumed for two-plate rheological model, the variation in

angular velocity of fluids under investigation in the parallel plate geometry from the centre of the

moving plate to the edge of the plate is another trade-off. The angular velocity increases linearly

as a function of the distance from the centre of the moving plate. Consequently, shear rate varies

across the plate. In order to make the shear rate independent of the distance from the centre of the

moving plate, one of the plates is substituted for a cone whose tip almost touches the other plate

at the centre. The angular velocity still increases linearly with the distance from the centre of the

moving plate, but now so does the gap between the plates, which in turn makes the shear rate

constant for the distance from the centre of the moving plate and therefore constant across the

plate. According to the ISO 3219:1994E, CEN standard, the cone angle, α is preferably close to

1°, and must not be greater than 4°.43 This geometry is an improvement on the parallel plate

because it requires smaller sample size and ensures homogenous deformation within the sample.44

The errors due to inertia and edge effects (where sulfur would be evaporating) are again significant

as with the cone and plate geometry. Therefore, it is also not suitable for accurate measurements

of rheometric properties of sulfur.

25
Concentric Cylinders (Cup and Bob)

The concentric cylinder rotational rheometer will be used for this study. In these rheometers,

materials are kept in the gap between a cylindrical cup and a bob. Maurice Couette in 1890 was

the first to make a practical rotational rheometer with concentric cylinder geometry.38 In Couette’s

design, the outer cylinder is made to rotate at a known speed while the inner stationary cylinder is

suspended by a torsion wire. A mirror was used to measure the angular deflection as a result of the

drag by the fluid on the inner cylinder. The angular deflection is indicative of the torque.

The second design of cup and bob geometry is the Searle geometry. In contrast to Coutte’s design,

Searle’s design has an outer fixed cup and an inner rotating bob to induce shearing, (Figure 6).

The shaft of the inner rotating bob also has attached to it the sensor of the torque required to sustain

the rotational motion.

ω
Inner cylinder

Ro
Sample space
L ω

Ri Outer cylinder

(a) (b)

Figure 6. (a) Concentric cylinder geometry of a rheometer showing its dimensions and the inner
bob; (b) cross-section of a concentric cylinder geometry of a rheometer.

26
By keeping the radii of the cylinders large and the gap between the cylinder inner wall and the bob

small, the difference in radial velocities at the two walls is small and negligible. Thus, the shear

rate between the walls is constant. In both Couette and Searle designs, the simple shear model is

assumed (Figure 3b). The inner bob is thought of as a folded upper plate and the outer cup as a

folded lower plate. The torque and rotational speed are the raw data obtained from rotational

rheometers. These data are combined with the physical dimensions of the measuring geometries

to obtain shear stress and shear rate, respectively, from which the viscosity is calculated. The shear

rate, γ̇ is obtained from equation (1-14), where Ro, is the radius of the outer cylinder, Ri, is the

radius of the inner bob, and ω is the angular velocity:

Ro 
  . (1-14)
Ro  Ri

If the torque is represented by M, then the shear stress, τ is obtained by equation (1-15),

M
  . (1-15)
2Ro2 L

It follows from equation (1-2), that viscosity, η measured at the inner shear rate is given by;

M ( Ro  Ri )
 . (1-16)
2Ro3L

1.8 Choice of Rheometer Geometry

Several factors are considered in determining the appropriate geometry to meet an experimental

need. Some of these criteria include the need to operate at high shear rate, uniform flow and

controlled temperature.45 Having described different rheometer designs, below is a summary of

their important of advantages and disadvantages:

27
Table 1. Advantages and disadvantages of different rheometers. (Adapted from Macosko, W. C.,
1994).38

Method Advantages Disadvantages

Capillary - High shear rate - Pressure drop correction is


- Wide viscosity range time consuming
- Sealed - Unsuitable for time dependence

Sliding Plates - Simple design - Shear rate limited to 10 s-1


- Homogenous flow - No gap control
- Best for high viscosity fluids - Difficult to load

Cone and Plate - Constant velocity gradient - Edge failure for high viscosity
- Induces homogenous flow fluids
- High inertia effects
- Evaporation of samples

Parallel Disks - Easy to load viscous samples - Non-homogenous


- Variable shear rate - Edge failure
- Evaporation

Concentric - Best for high shear measurements - Difficult to clean


Cylinders - Homogenous flow if cylinder ratio - Difficult to load with sample
( Ri / Ro ) ≥ 0.95

In addition to the advantages listed in Table 1, concentric cylinder rheometers possess some other

important characteristics that made it the most suitable geometry for the investigation of sulfur’s

viscosity. The concentric cylinder is the only rotational geometry in which samples are fully

confined to the annular region of the measuring profile to prevent undesired interaction from the

surrounding, evaporation and materials from being thrown out at high shear rate (Figure 6a).

Furthermore, this study aims at defining rheological parameters necessary to inject molten sulfur

at high shear rate into underground reservoirs for long term storage. It follows that a suitable

28
measuring configuration must be good for high shear rate measurements. The concentric circle is

unique for such measurements.

1.9 Rheology Models

Rheological data are oftentimes fitted to empirical or theoretical models. Equation (1-8) is a

constitutive equation that expresses viscosity as a material constant. In order to accommodate non-

Newtonian behaviour in materials, the exponent of the shear rate in the Newtonian constitutive

equation takes a value other than unity.46 The shear stress now depends on some power of the shear

rate, equation 1-18. This model is known as the power law or Ostwald-deWaele model:

  K c  m (1-18)

Where Kc, defined as consistency replaces viscosity in equation 1-18 and the exponent m is the

flow index. The value of m indicates the dependence of viscosity on shear rate. When

equation 1-18 is divided through by γ̇ , the power law equation takes the form in equation 1-19:

  K c  m1 (1-19)

A value of m ˂ 1 corresponds to shear thinning material behaviour and m ˃ 1 to shear thickening

material behaviour. When m = 1, equation 1-18 reverts back to the Newtonian constitutive equation

(equation 1-8).

Other variants of the power law model are the Sisko and Cross models.36,38 The Sisko model,

(equation 1-20) is an extension of the power law model with the addition of an upper limiting

Newtonian viscosity, η∞:

     K c  m1 (1-20)

29
For cross model, equation 1-21, reduced viscosity is used instead of viscosity to incorporate a

wider flow regime. Where η∞, is viscosity at upper shear limit and η0 is viscosity at lower shear

limit.

0 
 ( K c  ) m (1-21)
 

The Bingham model is a modified Newtonian (equation 1-8) constitutive law with the addition of

plastic yield stress for material behaviour that requires yield stress to flow, equation (1-22). In this

equation, yield stress, τ0 and plastic viscosity, ηp are empirical constants.

   0   p   (1-22)

30
Chapter Two: REVIEW OF LITERATURE DATA, EXPERIMENTAL TECHNIQUE
AND MODELS

Physical parameters such as temperature and shear rate have important effects on the flow

properties of liquid sulfur. Some important conditions include: (i) the freezing point of sulfur at

1 atm is approximately 115 °C, (ii) the viscosity of liquid sulfur abruptly increases to about four

orders of magnitude at T = 187 °C, and (iii) the auto-ignition of sulfur vapour at 3.5% mol (in air)

which will produce highly toxic SO2 gas must also be considered in the choice of experimental

technique for molten sulfur analyses. As a result, there are few experimental data for the viscosity

of sulfur. In this chapter, the available experimental techniques, data, as well as theoretical models

for sulfur’s viscosity will be reviewed. The deficiencies and strengths of these methods, the data

and models also will be highlighted.

2.1 Sulfur’s Viscosity Measurements

2.1.1 Measurements Data for the entire Liquid Temperature Range

Before Bacon & Fanelli reported their sulfur viscosity measurements in 1943 there had been two

data sources reporting the viscosity of sulfur without agreement. Rotinjanz (1908) used a modified

Ostwald viscometer (capillary viscometer) to measure the viscosity of sulfur in the entire liquid

range. The Ostwald viscometer is a U-tube which comprises of two bulbs separated by a capillary

section. The liquid sulfur is drawn into the upper bulb and the efflux time t, required for the liquid

to pass between two etched lines along the capillary tube while it drains into the lower bulb under

gravity is recorded. t is then multiplied by the apparatus constant K, to obtain; η = K t as explained

in subsection 1.7.4.1. Rotinjanz did not equilibrate the liquid sulfur before taking measurements

31
but carried out the experiment at heating rates ranging from 0.27 to 1 °C per minute.10,11,47 He

reported that η was dependent on the heating rate, with higher heating rate showing lower

maximum viscosity, ηmax values. His reported ηmax at 187 °C was 52,000 cP (52,000 units relative

to the η of water).2,10,48 It was also reported that the ηmax obtained was lower on cooling than it was

while heating the same sample. The latter are now understood to be due to reactions of impurities

in the liquid sulfur.

Later, Farr and Macleod (1928) found the modified Ostwald viscometer unsuitable for

measurements of liquid sulfur’s viscosity above 160 °C.10,11 They used a coutte-type rotating

cylinder viscometer for their measurements. In their set-up, they suspended a sealed glass tube

containing a mercury thermometer using a two-wire suspension system over a light pulley to

achieve equal tension across the rotating system.11 This glass tube was lowered into a test-tube

containing the liquid sulfur to be analyzed. The test-tube was suspended by a different one or two

suspension pulley system. The pulley was equipped with a motor with magnetic brake to vary the

speed of rotation. The rotation of the test-tube was followed by a mirror attached to the test-tube

from where the reflection of a spot of light on a circular scale was measured. The speed of rotation

was generally in the order of one revolution in five seconds and was recorded using a stop watch.11

A paraffin bath was used to maintain the temperature in the system up to 220 °C. The general

expression for pulley systems were combined with equations representing the rotation of cylinders

to calculate η, for the two suspension system is

( Ri  Ro )( Ri  Ro )m p gd1d 2
  sin θ , (2-1)
16Ri Ro hl
2 2

32
where Ri and Ro are the radii of the inner and outer cylinders respectively; d1 and d2 are the width

of suspension at the top and bottom respectively; h is the depth of immersion of glass tube in liquid

sulfur; l is the length of suspension; θ is the deflection angle; mp is the mass of suspended system.11

Just like Rotinjanz, Farr and Macleod had difficulties with the purity of sulfur. They observed that

the ηmax of sulfur was lower when cooling than heating and that a higher maximum heating

temperature gave lower ηmax value. Amongst many other observations, it was also concluded that

the ηmax of 21,500 cP of gas-free sulfur was reached at 200 ˚C. An air-exposed sulfur maintained

all night at 125 °C gave an ηmax of 42,750 cP while after exposure to air for a prolonged period at

about 190 °C, the viscosity of liquid sulfur could rise as high as 80,000 cP. They believed H2SO4

was the major impurity that greatly influenced the viscosity of air-exposed liquid sulfur.10,48

Bacon & Fanelli went the extra mile to purify the sulfur for their experiment by a rigorous

procedure they developed.6 These authors felt that most organic matter in liquid sulfur reacted

after a few hours of boiling, but the products of the decomposition reactions appeared to be

resistant to further destruction by heating (now referred to as CarSul production). They subjected

the commercial grade sulfur (Frasch) to a prolonged cycles of heating, boiling, solidification,

melting, crystallisation and vacuum treatment in the presence of magnesium oxide. The

magnesium oxide reacts with acid impurities and catalysed further decomposition of H2Sx present

in liquid sulfur. The liquid sulfur was then filtered and tested for organic matter to confirm it was

free of these impurities. Four cycles of heating for thirty hours were suggested to achieve complete

elimination of carbonaceous matter, H2S and H2Sx in liquid sulfur.6

33
For their measurements, Bacon & Fanelli used a modified Ostwald viscometer for the temperature

range between 118 °C and 160 °C and a paraffin bath for temperature control. For measurements

above 160 °C, they used a pipet graduated in milliliters. A known vacuum pressure was impressed

on the capillary tube immersed in the liquid sulfur to be measured. The time taken for the sulfur to

rise to a marked length of the capillary was recorded using an electric stop clock. This set-up was

housed in a glass wool lagged electric air-bath for temperature control. The viscosity was

calculated using the Poiseuille’s equation,

r 4 pt .
 (2-2)
8VL

From this measurement, ηmax of liquid sulfur was reported to be 93,200 cP at 187 °C. They also

treated different portions of their purified sulfur under vacuum and equilibrated at different

temperatures for different durations. After several measurements, they concluded that the rate of

heating, cooling and thermal history have no effect on the viscosity of sulfur; and from their data

the heating curve of η vs. T measurements of pure sulfur coincided with the cooling curve. They

also assigned the discrepancies in previous literature data to the impurities present in sulfur used

for the experiments. Doi (1963) also made measurements using a rolling ball viscometer. His data

are in agreement with the experimental data of Bacon & Fanelli. (Figure 7).49

34
Bacon & Fanelli
λ-transition
50000

5000
Farr & Macleod
Sn ⇌ •S•n
ɳ / cP

500

50
Ruiz-Garcia et. al
Doi

5
120 160 200 240 280 320
T / °C

Figure 7. Experimental literature data of sulfur’s viscosity. ▲, data of Farr & Macleod;48 ●, data
of Doi;49 ♦, data of Ruiz-Garcia et al.;29 ■, the heating profile data of Bacon & Fanelli; 2 ■, cooling
profile data of Bacon & Fanelli.2

2.1.2 Measurements Data below 160 °C

Ruiz-Garcia et al. (1989) used a falling ball viscometer for their measurements. A high-purity

sulfur sample (99.999%) was degassed by melting, followed by the evacuation of evolved gasses

for fourteen cycles.29 The degassed sample was sealed into a glass tube containing glass a ball and

having two reference marks. A stopwatch was used for timing the fall time of the glass ball between

the reference marks. Stokes’ law was used for the calculation of the viscosity:

35
  K  B  l   t , (2-3)

where K is the instrument constant, ρB the density of the glass ball and ρs is the density of sulfur.

Within the temperature of interest (118 °C - 163 °C), ρ of sulfur changes by 1 %, while the density

of the glass ball changes by 0.02 %.18,29 Ruiz-Garcia et al. observed a reduction of viscosity values

i.e., relaxation when measurements were collected continualy for three days. These viscosity data

were still larger than those reported by earlier authors. When data were collected within three

hours, not more than 0.5 – 1 % relaxation in the viscosities were observed. The reported data of

Ruiz-Garcia et al. corresponded to sulfur being heated versus cooled between experiments. Like

previous authors, they concluded that the relaxation was due to a trace amount of impurities present

in liquid sulfur. Like many other authors, they reported only the highest viscosities obtained at

each T during their measurements.2,10,49 Their data appear to be the better than all earlier data

reported considering the higher purity of the sulfur for their experiment. Figure 7 shows the

experimental data of the workers whose methods have been discussed above. The data of Farr and

Macleod, Bacon & Fanelli, and Doi are within (1 - 2% ) of one another for the measurements

between the sulfur melting point and T = 160 °C.

Apart from the response of the timer being a limitation of the method of Ruiz-Garcia et al., they

also acknowledged that liquid sulfur might possess a non-Newtonian behaviour especially at the

higher ranges of T and η. This they claimed can have an insignificant effect on sulfur’s viscosity

within the temperature of their interest which is between melting and 160 °C.29 They also claimed

that the shear rate for their measurements must have varied only between 6 s-1 to 60 s-1. Previous

authors neither mentioned nor accounted for the shear effect on the η data of liquid sulfur they

reported. It is therefore important to revalidate the viscosity data especially for temperature above

36
160 °C, knowing that previously reported measurements might have been influenced by shear rate

which was not accounted for.

2.2 Viscosity Models for Sulfur

Because of the wide industrial applications of sulfur, there is need for accurate estimation of its

viscosity among other physical properties. This will help in optimizing the Claus plant operations;

and processes involving sulfur pipelines, pumps, tankers and storage tanks. To define the region

with a sharp rise in viscosity after the onset of λ-transition of sulfur, a number of authors have

developed several models based on the theoretical principles discussed below.

2.2.1 Equilibrium Polymerisation Theory

Powell & Eyring (1944), Gee (1952), Tobolsky & Eisenberg (1958), and Touro & Wiewioroski

(1966) have all worked on modelling sulfur’s viscosity based on the theory of equilibrium

polymerisation.50,51,52 The objective of these studies was to quantify the polymer - monomer

equilibria of molten sulfur and in turn correlate η of sulfur to its equilibrium polymer concentration.

Powell & Eyring assumed that the net reaction taking place was breaking the rings which leads to

the formation of radicals of different chain lengths. And thus was represented by the reaction

x
S8 (ring) ⇌ • Sx• (chain) where 0 < x < ∞. (2-4)
8

Assuming a constant enthalpy of reaction ∆H, the equilibrium constant Kx is given by

yx
Kx  x /8
 e  H / RT  ( as x bs ) / R , (2-5)
(y )

where R is the ideal gas constant, x is the chain length, yx is the mole fraction of molecules with

length x and y‫ ٭‬is the mole fraction of S8 rings. as and bs are entropy terms due to end effects and

37
chain lengths respectively. They calculated the mole fraction of the molecules in • Sx• to that in S8

rings and in turn the weight fraction, ϕ of polymer to monomer in liquid sulfur as a function of T.

This calculation was compared to the sulfur polymer equilibrium experimental data of Hammick,

Cousins and Langford obtained by quenching liquid sulfur rapidly to room temperature before

extraction with CS2 to find its equilibrium compositions.50,51,52 There was a good agreement where

the experimental data were reliable.50 Two further assumptions were made to correlate the

calculated ϕ to the viscosity of sulfur;

(1) The η of sulfur can be represented by the Flory equation for molten polyester, relating η to

average chain length and molecular weight; η α (𝑟̅ 2)3/2 / Mav , where 𝑟̅ is the

root-mean-square end-to-end average chain length of a polymer and Mav is the average

molecular weight of the polymer,

(2) The general η equation for a polymer mixture holds for the • Sx• and S8 mixture;

M  (rings)1 (chains) (2-5)

Calculating η from a sulfur-halogen system resulted in the following expression:

1/ 2  
ln = ln  M 
BZ w 1 1  (2-6)
R  1 M x Z 
 

Where ηM is viscosity of mixture, η is viscosity of sulfur, Mx is the ratio of halogen atoms to S

atoms and Zw is the weight average change length. B and Z are constants calculated from other

available literature and theoretical data. Even though the shape of the plot of this model nearly

matched the data of Bacon & Fanelli, it over-predicted the maximum viscosity of sulfur by at least

a factor of two.51

38
Gee (1952) pointed out that the theoretical model of Powell & Eyring did not account for the

critical temperature at which a sharp rise in viscosity occurred.50 He then proposed an equilibrium

based on removal of S8 rings from the ends of a diradical chains as



Sx• ⇌ • S •x-8 +S8 (2-7)

By applying Flory’s law of polymer mixtures and neglecting the T dependence of ρ, ϕ of polymer

in equilibrium was obtained as

 H  1 1  
 1 exp      . (2-8)
 R  TΦ T  

According to this expression there is a discontinuity at Tϕ (transition temperature), and ϕ will be

stable below or above Tϕ depending on whether or not the ∆H is negative or positive.50 This made

the equation invalid below Tϕ, even though the hypothesis that the production of polymer started

at a critical temperature was satisfied. Because this theory could not satisfactorily predict the ϕ

below Tϕ and at T very near Tϕ, Gee applied a different theoretical approach to predict the ϕ below

Tϕ. He used the Arrhenius equation, which fit well to sulfur’s η experimental data between melting

at 115°C and 160 °C to account for the low values of η below Tϕ. Using the experiment data of

Hammick, Cousins and Langford, a plot of -ln (1 – ɸ) against 1/T gave a satisfactory straight line,

from which Tϕ and ∆H4 / R were obtained to be 150 °C and 16,736 K respectively. A general

expression for η dependence on chain length was used to account for sulfur’s polymer chain

length.50

That is,

2
  AP 3 , (2-9)

where A is a constant which depends on T, and P is the average polymer chain length, and for

polymer concentration, equation (2-10) was used,

39

 0 1  kT  ,
2 2
(2-10)

where kT is another constant that vary as T, Tϕ was adjusted from the calculated 150 to 159 °C to

fit the region of rapid η increase. Gee’s final expression for the dependence of η on T was obtained

as

 2 ΔH 5  1 8   
ln η=B   2940+  + ln 1- exp ΔH 4  1  1   . (2-11)
 3R  T 3  R  Tφ T  

In this expression, three unspecified constants were combined into B, then B and ∆H5 were used

as fitting parameters. The predictions from this model matched experimental data well except for

the narrow T range around Tϕ where the predictions are poor.

Tobolsky & Eisenberg proposed a unified theoretical treatment to account for sulfur’s ϕ below and

above Tϕ by using two equilibrium reactions to describe the polymerisation in liquid sulfur. The

first reaction was ring opening while the second was polymer propagation reaction:

S8 ⇌ • S8• (2-12)

S8• + x S8 ⇌ • Sx• (2-13)

Tobolsky & Eisenberg’s equilibrium concentrations were calculated in molality (m; mol kg-1) to

avoid the use of ρ, whose dependence on T would have impacted uncertainties into the predictions

from Gee’s model.50,52 In agreement with Gee, ∆H4 / R from Tobolsky & Eisenberg’s theory was

obtained to be 16,520 K for reaction 2-12 and 15,960 K for reaction 2-13.52 Moreover, Eisenberg

(1969), using Fox and Allen’s extended viscosity theory based on molecular flow, found the power

law index of 3.4 to be operative for sulfur polymer entanglement system. Combined with the above

unified theory of polymer chain length, concentration and using the modified William-Landel-

40
Ferry (WLF) empirical equation for monomeric friction factor, η was calculated. ηmax was four

orders of magnitude higher than that of the experimental data.53 Eisenberg then suggested that

theory based on molecular flow alone could not accurately predict the η of sulfur.53

2.2.2 Modified Reptation Theory

Though Eisenberg developed further equations based on bond interchange and chain-end

interchange theories, both over predicted η of sulfur by four orders of magnitude.53 Cates (1987)

incorporated bond breaking into the reptation theory, combined with existing sulfur polymer

equilibria theories established by earlier authors. Reptation motion is the thermal motion of long

linear entangled polymers in polymer melts or concentrated polymers, analogous to snakes

slithering through one another.54 A polymer chain is restricted to motion along a tube of

entanglements made by other polymers. By exiting the tube in a snake-like diffusion, relaxation

takes place in a time scale known as reptation time (τrep). τbreak represents the time taking for a

chain of polymerisation index N, to break into two pieces. For τrep ˂˂ τbreak, the rate-limiting step

in relaxation of stress associated with a particular portion of the chain would be the waiting time

for a break to appear within a certain distance of chain length. In modifying this motion to

accommodate chain breakage and recombination, the following assumptions were made:

(1) Recombination of chains is typically at the end of another chain, rather than the one

that immediately dissociated from the chain

(2) Breaks can occur anywhere along the • Sx• with equal probability per unit time

(3) Addition and breaks of S8 monomers were disregarded in finding the lifetime of chains

Assuming breaks can occur with equal probability per time, the Flory expression for polymer chain

conformation is used to obtain stress relaxation time as

41
Nl 2
 rep  break
r   ,
2
break
(2-14)

where l, is a monomeric chain length and r2 (t) is the mean-square displacement of a chain end in

time t in reptation model.38 For this time scale, the η obeys

  Ge , (2-15)

where Ge is the plateau (equilibrium) modulus of the entangled polymer system. For the liquid

sulfur system, Cates reasoned that the recombination time, τbreak is so short that the motion of chain

end is on the time scale of its recombination is always Rouse-like.55 Therefore,

N
   break , (2-16)
n

 f n 2   Rouse(n)   break , (2-17)

where, τf is the friction constant and n is the number of monomers. To calculate η; N, Ge, τf and

τbreak must first be obtained as functions of T. N and the volume fraction (ϕv) were obtained from

the equilibrium theories of Gee;

  1 1 
v 1 exp 11610    , and (2-18)
  4321 T  

ln N    4.14  ln  v  .
7550
(2-19)
T

Ge, the plateau modulus was found as

cRT  N 
Ge  1  2 e  , (2-20)
Ne  N

where c is the number density proportional to ϕ and Ne is the polymer entanglement length.55,56

The temperature dependence of τf was obtained using the WLF empirical equation, with the

parameters of Eisenberg (1968),53 as

42
1510 26.2T  243
ln  f  C  
T 195
, (2-21)
T

The expression for τbreak was found to be

 break  Nk 1 , (2-22)

where k is the first-order rate constant for the dissociation of S-S bonds in polymerised sulfur.

Bacon & Fanelli’s data were fit to a modified Arrhenius equation in the form: By combining

equations 2-14 to 2-22, the expression for η was found to be;


1
 12
  Ge N  f k
1
2 2 . (2-23)

Bacon & Fanelli’s data were fitted to a modified Arrhenius equation in the form;

k/λ2 = A exp (-B/T), while A & B, along with C in equation 2-21 were used as fitting parameters. λ

is a constant of the order of unity used to absorb possible errors in equation 2-20.55 The following

equation was therefore obtained by Cates

k  15400 
 7.11013 exp  , (2-24)
 2
 T 

where B (15,400 K) is the activation energy for the dissociation in the form Ea / R, and λ is a

product of constants independent on T. Only A and B in the expression for k were used as fitting

parameters as k from previous works are not in agreement.45,52,55 The activation energy, B was

obtained as 15,400 K which differs from the value 16,500 K of both Gee and Tobolsky &

Eisenberg.45,52,55,57 This was attributed to the uncertainties in the WLF parameter’s

pseudo-Arrhenius behaviour at T > Tɸ, which was confirmed by extrapolating the model to a lower

T range to obtain k values that are within the range of estimates of previous works. However, this

model fit well to the experimental data of Bacon & Fanelli (1943).6 Cates calculated τrep for a

hypothetical, unbreakable and entangled polymer from this model, to be 1500 s. This is five orders

43
of magnitude larger than τrep of 1.5 x 10-2 s obtained for sulfur polymers using this model. Again,

this confirmed the hypothesis by Eisenberg that models based on the molecular flow of sulfur

alone would overestimate the η of sulfur up to four or five orders of magnitudes.

Being the only model that incorporated bond breakage into the polymerisation theory, this model

far better represents the liquid sulfur equilibria compared to earlier efforts. Again, the verifiable

assumptions made for this model were validated. The sulfur viscosity data of this thesis will be

fitted to the modified reptation model of Cates (equations 2-23 and 2-24).

2.3 Effects of Impurities on Sulfur’s Viscosity

The following impurities: H2SO4, H2S, amines, halogens and hydrocarbons have been shown to

have a modifying effect on the viscosity of molten sulfur.10,58 Of these impurities, H2S is of most

importance for reasons mentioned in section 1.6, and so, its modifying effect on the viscosity of

sulfur will be reviewed in this thesis.

2.3.1 Effects of H2S on Sulfur’s Viscosity

Fanelli (1946) studied the modifying effects of H2S on the viscosity of molten sulfur and later, in

a separate study, measured the solubility of H2S in molten sulfur.10,12 For the determinations of the

viscosity modifying effects, a known amount of H2S was introduced into a specialized capillary

viscometer in the form of liquid H2S or sulfur disulfide (H2S2) before viscosity measurements were

made. In the latter, H2S2 readily decomposed into H2S. The H2S content was calculated from the

H2S2 added as weight percent of the total weight of mixture. It was observed that much of the H2S

remained in the vapour space during the η determination. After saturation with H2S, while the

44
capillary viscometer was being sealed before viscosity determination, it was reported that some

H2S gas escaped. This caused underestimation of the modifying effects of the dissolved H 2S on

the viscosity of sulfur.10

Rubero (1964) used a two autoclave capillary-viscometer-system to re-examine the solubility of

H2S in molten sulfur.58 He completely filled the first autoclave with molten sulfur and subsequently

with H2S as a portion of the molten sulfur was drained out. The second autoclave, which was

connected to the first by a metering tube, was filled with nitrogen gas. The two autoclaves were

connected to a mercury barometer from which the pressure was directly recorded. By using the

Poiseulle’s equation (equation 2-2), the viscosity was calculated from the pressure difference (∆p)

and the instrument constant (K). To quantify the dissolved H2S, the molten sulfur was allowed to

flash into a tared container in the second autoclave which was connected to an iodine filled

absorption train. The excess H2S in the absorption train was obtained by titrating with thiosulfate

and the weight of sample was obtained by reweighing the tared container after the experiment. As

expected, Rubero recorded a higher suppression of sulfur’s viscosity than Bacon & Fanelli for the

same amount of dissolved H2S.58 However, elemental mercury (Hg) reacts with sulfur to produce

cinnabar (HgS) which is a very stable compound. There is a high likelihood of occurrence of this

reaction as sulfur vapour was in contact with mercury in the barometers used for pressure

measurements. Hence, this method is inherently defective.

Touro & Wiewiorowski (1965) showed that pure molten sulfur is absorption free for infra-red (IR)

vibration spectroscopy between 1400 to 4000 cm-1 and so, can serve as a solvent for the IR

determination of dissolved H2S in sulfur.59,60 They scanned molten sulfur treated with H2S in the

45
region of S-H stretching (2400 – 2600 cm-1) where absorption peaks at 2498 cm-1 and 2570 cm-1

were observed.59 They assigned the former to H2Sx and the later to H2S after analyses of the rate

at which each peak developed.

Finally, Marriott et. al. (2008), used a more sensitive IR vibration spectroscopic technique to

measure the solubility of H2S (H2S+ H2Sx) in liquid sulfur, and to further characterize the

physically adsorbed species (H2S) from those chemically dissolved H2S (H2Sx) at different T and

p corresponding to conditions encountered during the industrial processing of molten sulfur.24

They observed the asymmetric vibrational IR absorption of the S-H bond for H2S in liquid sulfur

at a wavenumber of 2571 cm-1, while the S-H bond for H2Sx was observed at 2498 cm-1.24 The

ratios of H2Sx : H2S in liquid sulfur obtained from this method were different from those reported

by Touro & Wiewiorowski.24,59,60 Note that the deficiencies highlighted above in the experimental

solubility data of Bacon & Fanelli must have imparted the H2Sx : H2S ratio reported by

Touro & Wiewiorowski as the latter’s T correction factors were derived from the former’s

experimental data. In addition, the long formation time necessary for H2Sx and in turn possible

absorption from unterminated HSx• radical species which also must have imparted the experimental

data of Touro & Wiewiorowski likely contributed to the discrepancy. The IR technique of

Marriott et. al. will be used in this study.24

46
Chapter Three: EXPERIMENTAL METHOD AND PROCEDURE

An Anton-Paar Modular Compact rheometer (MCR 302) rated to 1000 bar was purchased for this

study. To commission this instrument, it will first be used to analyze available certified viscosity

and rheology standards. Unique procedures must be established to safely deliver molten sulfur into

the pressure cell while accurately measuring the head pressure and the rheometric properties of

liquid elemental sulfur.

3.1 LabVIEW Code for Pressure Data Acquisition

LabVIEW code (Laboratory Virtual Instrument Engineering Workbench) was written to log

pressure measurements from a Keller PA-33X transducer (0 - 1000 bar) to a computer file. The

LabVIEW environment is different from text based programming languages (such as Fortran

and C), in that LabVIEW uses a graphical programming language known as the G programming

language to create programs relying on graphic symbols to describe programming actions.61,62

LabVIEW has two interfaces that are connected to each other. The front panel which has the

controls and the indicators, and the block diagram which contains the actual executable program.

Once the Virtual Instrument (VI) is made and running, a user may control or give instructions to

the VI through the front panel but cannot interfere with the block diagram until the program is

stopped.

Figure 8 shows the front panel of the VI created for this study while its block diagram can be found

in the Appendix A. In this front panel, two numeric controls B and C were created: B is used to

key in the respective file name, while C is used to select the communications port of the transducer

47
on the computer (COM 5). In addition, separate numeric controls were created to input the

calibration constant and the vacuum correction constant into the VI.

A F

H
I

A
E

Figure 8. Front panel of the VI to log pressure from the transducer. Notations: A, pressure readings
indicator; B, numeric control for file name; C, communication port selector; D, boolean for data
saving; F, pressure reading graphical display ; G, transducer communication indicator; H,
calibration constant input numeric control; I, vacuum correction input numeric control.

Two indicators A and F were drawn: “A” displays the real time numeric pressure measurements

while “F” is a graphical display of the pressure measurements history. Boolean was created: one

was to give the command whether or not to save data (D), while the other was to indicate whether

or not the transducer was communicating with the program (G). A while loop was drawn to give

conditions for the continuous running of the VI (E); however, the loop and program could be closed

by pressing the “Stop” switch. Numeric controls H and I were also created to input the calibration

48
constant and vacuum corrections respectively. The wirings between controls and indicators of the

VI can be found in the corresponding block diagram in Appendix A.

3.2 Pressure Transducer Calibration

3.2.1 Deadweight Tester

A deadweight tester (DWT) or piston gauge uses a known mass, m, of a set of weights to generate

a force, F, over a cross-sectional area, Acs. The pressure p, is then counterbalanced by exerting an

equal pressure from a pressure pump on the fluid in the DWT reservoir until the weights freely

floats indicating an equilibrium of pressure throughout the hydraulic system.62,63 By connection to

the fittings on the DWT, the p generated is transferred to the transducer to be calibrated. p is given

by:

F  mi g
p  i (3-1)
Acs Acs

To ensure accurate measurements, the transferred p, is corrected for local gravity, gl, local

temperature, buoyancy of the stack of weights, thermal expansion of the instrument and static

pressure head as a result of the different height between the transducer and the piston. With the

incorporation of the corrections, equations 3-1 is transformed into:

 mi g   
1  air   g l  fluid h 
F
p  i
Ap Ap   mass 
, (3-2)

where

 
Ap  Ae 1   p   c T  Tref  , (3-3)

and

49

Ae  a0 1  bd p . (3-4)

a0 represents the effective area of the piston at ambient pressure and the reference temperature,

Tref, at which the properties of the deadweight piston were initially measured. The pressure

distortion coefficient, bd, in equation 3-4 is used to calculate the constant temperature effective

area, Ae, when a certain amount of pressure is loaded. Ae is then converted to the temperature-

corrected true effective area, Ap, through equation 3-3 for the temperature T, at the time of

calibration, using the coefficients of thermal expansion for the piston and cylinder, αp and αc,

respectively.62, 64 Once Ap has been determined, the buoyancy correction is applied to the pressure

equation using the ratio of air density, ρair, to the density of the weights, ρmass.62,64 The static

pressure head correction is also applied due to the height difference, Δh, between the transducer

and the bottom of the piston assembly using the local acceleration due to gravity, gl, and the density

of the hydraulic fluid, ρfluid.

3.2.2 Experimental Methods

A Pressurements Limited T 3800/4 Deadweight Apparatus and standard weights were used to

apply the reference equilibrium pressure to the Keller pressure transducer. The standard weights

were carefully cleaned and measured on a highly sensitive X26003L Mettler-Toledo mass

comparator to 3 decimal digits (0.001 g). Isopropyl alcohol was used as the hydraulic fluid instead

of a commercial hydraulic oil for ease of cleaning the transducer’s measurement chamber after the

calibration.

To generate pressures for the calibration, the following procedures were used (Figure 9). The

leveling feet J of the dead weight tester were adjusted to ensure balance while observing the bubble

50
level in the indicator “G”. The pressure valve “F” was opened and the capstan “D’’ (Figure 9a)

was turned fully out. With the use of the hand pump the hydraulic fluid transfer lines were pre-

pressurized and de-aerated to prime the system. Then, weights equivalent to the desired pressure

were placed on the weight stack “H” and the hand pump “A” was used to pressurize the system

until the desired pressure was nearly reached. The capstan was screwed-in until the weight stack

floated (Figure 9b). The weights were adjusted until the underside edge of the bottom weight was

on the mid reference mark of the indicator rod “I” (Figure 9a) to ensure proper alignment of the

weights. The weight stack was then spun in the clockwise direction between 40 to 100 rpm so as

to obtain a consistent pressure. The spinning helped to minimize error that may result from static

friction between the piston and the cylinder.63,64 At this point, the pressure readings on the

LabVIEW graphical indicator were observed to be constant, then the program was prompted to

save the pressure readings.

51
(a) (b)

Figure 9. (a) DWT without weight on the weight stack. (b) DWT with transducer and weight on
stack at equilibrium with pressurized fluid. Notations: A, de-aeration hand pump; B, pre-
pressurization hand pump; C, hydraulic fluid reservoir; D, fine-pressurization capstan; E,
connection fitting for pressure device to be calibrated; F, pressure-release valve; G, bubble level;
H, weight stack holder with base plate in place; I, base plate indication rod; J, adjustable foot for
levelling; K, Keller pressure transducer; L, floating loaded standard weight.

3.2.3 Results and Discussion

Correction parameters for gravity, buoyancy and pressure head unique to this instrument as

provided by the manufacturer were summarized in Table 2. Equations 3-2 through to 3-4 were

provided also by the manufacturer to calculate the theoretical pressure exerted by the associated

masses. Also, the required masses were obtained by the addition of all the weights used in exerting

the pressure. The corrected theoretical pressure from equation 3-1 and the respective pressure

measured by the pressure transducer are given in Table 3.

52
Table 2. Relevant parameters for deadweight calibration.64

Parameter Value

gl 9.8082 m·s-2
a0 4.06115 × 10-6 m2
bd 1.043 × 10-8 MPa-1
αp+αc 2 × 10-5 °C-1
T 21 °C
Tref 20 °C
ρair 1.22 kg·m-3
ρmass 7300 kg·m-3
ρfluid 785 kg·m-3
Δh 0.0635 m
pvac,21°C 0.0121 MPa

Table 3. The corrected theoretical pressures, the measured pressures, and uncertainties of the
measured pressures estimated using the standard deviation and a 95.5 % confidence intervals.

Ptheoretical / MPa Pmeasured / MPa δ / MPa

4.000 3.996 0.019


5.000 4.997 0.014
6.000 5.996 0.015
7.000 6.994 0.017
8.000 7.995 0.011
9.000 8.996 0.006
10.000 9.995 0.017
12.000 11.993 0.017
14.000 13.995 0.007

53
Table 3. Continued

Ptheoretical / MPa Pmeasured / MPa δ / MPa

15.000 14.994 0.010


18.000 17.993 0.012
20.000 19.992 0.016
22.000 21.991 0.018
22.997 22.983 0.015
24.997 24.983 0.017
29.996 29.982 0.011
31.996 31.980 0.020
39.995 39.976 0.020

The corrected, calculated pressures were plotted against the pressure measurements made using

the pressure transducer. The measured pressure also was corrected for atmosphere by measuring

and subtracting the temperature-dependent vacuum pressure reading. A linear regression was

completed on the plot to obtain the calibration and is shown in Figure 10. The resulting calibration

equation is given as:

Pcorr  1.000468 0.000036 Pmeas . (3-5)

54
45

40

35

30
pcorr / MPa

25

20

15

10

0
0 5 10 15 20 25 30 35 40 45
pmeas / MPa

Figure 10. Deadweight calibration plot of the calculated applied reference pressure (pcorr) against
their corresponding transducer measurements (pmeas).

Because errors in Table 3 (95 % confidence interval) were all greater than the error associated with

the calibration coefficient (0.0036%), the quoted error for each pressure measurement will

therefore be estimated using the standard deviation and the 95 % confidence interval.

3.3 Commissioning the MCR 302 Rheometer

The MCR302 is a modular rheometer in which different measuring geometry and temperature

devices could be coupled. The pressure cell geometry is the only suitable measuring system for

this study because for other geometries, sulfur vapour will continuously escape which would

55
impact experimental data and constitute a safety concern. Other factors that make other rheometry

measuring geometries unsuitable for this study have been discussed in chapter two.

3.3.1 Operation and Hardware of MCR 302 Rheometer

The instrument is equipped with an electric DC motor supported by two air bearings, where the

current supplied to the stator coils is proportional to the torque of the motor.65 The radial air bearing

centers and stabilises the shaft while the axial air bearing holds the weight of the rotating parts.

This instrument can operate with a torque (M) range from 1 nNm to 2 x 108 nNm.65 A magnetic

coupling, attached to the shaft of the motor, imparts a rotational motion on the sample sandwiched

between the pressure cup and a cylindrical magnet by rotating the cylindrical magnet (Figure 11).

The cylinderical magnet is suspended between two sapphire bearings which minimize the friction

at the contact points of the cylindrical magnet with the pressure cup. The pressure cell and the

cylindrical magnet are made of titanium for corrosion resistance. Heating of the pressure cell is

done resistively and is rated up to 300 °C. A 100 ohm platinum resistant thermometer (PRT) is

embedded in the cell for temperature measurement.

56
MCR 302 rheometer
Keller pressure Sulfur
transducer reservoir Electric
motor
Magnetic
coupling
Rotating
V1 cylinder
Sample
V2 V3 space
Pressure
High-pressure

cell
N2 cylinder

Sapphire
bearing

V4

Thermostated FT-IR cell

Figure 11. Schematic of the experimental set-up for measurements of the rheometric properties of
liquid elemental sulfur.

57
Figure 12. Picture of the experimental set-up for measurements of the rheometric properties of
liquid elemental sulfur.

3.3.2 Controls and Data Logging

Rheoplus 4.14 is the software from the manufacturer (Anton Paar) of the rheometer used to send

instructions to the instrument and log measurements from it to a computer system. Using this

software, the instrument can be operated in controlled shear (CS) mode - where the torque (M) is

set and the angular speed (ω) that is obtained as a result of the drag by the sample on the cylindrical

magnet is measured to obtain η, or the controlled rate (CR) modes - where the ω is set while the

58
M required by the electric motor to maintain the ω is measured to obtain the η. The temperature,

time, shear stress, shear rate, speed, and torque can all be varied or kept constant to obtain different

information from sample analyses (Figure 13).

Figure 13. User interface of Rheoplus 4.14 software used for controls and logging rheological data
from MCR 302 to the computer system showing some input variables.

3.3.3 Comparing Measured Results with Standards

To check the accuracy of the MCR 302 rheometer, two Newtonian standards and a non-Newtonian

standard were analyzed. The Newtonian standards included: a low viscosity standard – (Canon

Instrument Company N600) with a viscosity range from 1031 cP at 25 °C to 16.61cP at 135 °C

and a high viscosity standard - (Canon Instrument Company N450000) with a viscosity range from

1,539,000 cP at 25 °C to 2,509 cP at 135 °C. These standards are traceable to American Society

59
for Testing and Materials (ASTM) and National Institute of Standard and technology (NIST)

methods.66,67 The Newtonian standard was Standard Reference Material 2490 purchased from

NIST.68 The certified viscosity values for this standard are provided in Table 4.

Table 4. Certified values for NIST rheology standard-SRM 2490.

T / °C γ̇ / s-1 η / cP δ / cP

25 6.31 41440 270


25 10.00 33040 190
25 15.85 25610 140
25 25.12 19360 100
25 39.81 14260 80
25 63.10 10220 70
25 100.0 7220 60
50 6.31 23540 370
50 10.00 19950 320
50 15.85 16410 270
50 25.12 13140 220
50 39.81 10240 170
50 63.10 7750 130
50 100.0 5720 100

3.3.3.1 Experimental Method

The standard samples were stored and handled as suggested by the manufacturers to obtain

consistent results. To load these samples, the procedure described later in section 3.7 was followed

with the set-up temperature set-point (SP) set to the minimum temperature of analysis for each

60
standard respectively before loading them. Each standard was loaded and measured three times at

the temperatures and shear rate specified in the respective test certificates.66-68

3.3.3.2 Results and Discussions

The measurements made for the Newtonian standards are plotted in Figure 14. The comparison

between the measurements of N450000 and its certified values shows measurements within 10%

of the certified values at 101,500 cP, and within 3% at 2350 cP. Also for this standard, a higher

confidence interval of 99.7 % were calculated using the t-distribution and plotted instead of the

standard 95.5 % confidence interval.69 This is because the 95.5 % was difficult to observe on the

graph. The certified values are well within the plotted 99.7 % confidence interval of the

measurements. For N600, the measurement data were within 1 % of the certified values at 1031

cP and 2 % at 164 cP. A plot of 95.5 % confidence intervals shows the mean of the triplicate

measurements was in agreement with the standard (within the confidence interval) at η = 1031 cP

and 164 cP. While the measurement at η = 16 cP was just outside the 95.5 % confidence interval.

This viscosity value is however outside the range desired for this study and is not of concern.69

61
100000

10000

1000
ɳ / cP

100

10

1
20 40 60 80 100 120 140
T / °C

Figure 14. Comparison of viscosity measurements made using the MCR 302 rheometer and
certified values. ▲, certified values of N600; ∆, measurements of N600; ●, certified values of
N450000; and ○, measurements of N450000. Error bars on the experimental data are at a
confidence interval of 99.7 % for N45000 and 95.5% for N600.

The measurements of the non-Newtonian standard are shown in Figure 15. For these

measurements, the uncertainties ranged from 0.02 % at 6 s-1, to 9% at 100 s-1 of the certified values.

Again, the error bars are calculated with a higher confidence interval of 99.7 % using the

t-distribution instead of the standard 95.5 % confidence interval because the 95.5 % was hard to

observe on the graph.69 The certified values are within the 99.7 % confidence interval for each set

of measurements across the measured shear rate range.

62
50000

40000

30000
ɳ / cP

20000

10000

0
0 20 40 60 80 100 120
-1
γ̇ / s

Figure 15. Measurements of NIST’s SRM 2490: ●, certified values at 25 °C; ○, measurements at
25 °C; ▲, certified values at 50 °C; and ∆, measurements at 50 °C. Error bars on the experimental
data are at a confidence interval of 99.7 %.

3.4 Experimental Set-up

A negative-pressure bay equipped with automatic toxic gas detector was made available for these

studies. The set-up in the bay (Figure 12) comprises of a HP (high-pressure) nitrogen cylinder well

secured to the wall with steel chains. One-sixteenth inch tubing connects the high pressure cylinder

valve (V1) to the Keller pressure transducer. One-quarter inch HP tubing connects the transducer

into the first inlet port of a two-way valve (V2). The molten sulfur reservoir is a 100 mL, 316L

stainless steel cylinder, which fits into the second of the two inlet ports on the two-way valve (V3).

The dual valve outlet, through one-quarter inch HP tubing is connected to the tee-fitting under the

MCR 302 rheometer. With HP quarter-inch tubing, the upper port of the tee directly connects to
63
the HP cell, while the lower port connects to a vent/waste valve (V4). As shown in Figure 11 and

12, the molten sulfur reservoir, the HP tubing and the valves were all heat traced and insulated for

temperature control. All dimensions of tubing used were made of the 316L alloy. The temperature

controllers use the PID (Proportional Integral Differential) feedback-loop control mechanism to

continuously minimize the difference between the Process Variable (PV) which is the actual

temperature, and the set point (SP) which is the desired temperature. The SP is set to ca.140 °C to

ensure that the temperature in the molten sulfur reservoir, HP tubings, and valves, do not rise to

160 °C or fall below the melting point of sulfur at approximately 120 °C.

3.5 Materials

Oxygen-free, nitrogen gas (99.998% as per certificate of analysis) was purchased from Praxair

Technology, Inc. ACS grade, 99.5% toluene was supplied by BHD through VWR Analytical. The

sulfur prills used for this study were supplied by Keyera Energy. The total analysis of these sulfur

prills was performed by Alberta Sulfur Research Limited as shown in Table 5.

Table 5. Result of the total analysis of Keyera air prilled sulfur.70

Moisture content Purity Ash Content H .C. Content H2S Content


Sample
(wt %) (wt. %) (ppmw) (ppmw) (ppmw)

A 0.02 99.992 4 0.6 0.4

B 0.09 99.991 14 0.6 0.3

64
3.6 Experimental Procedure

To prepare the apparatus for the measurements of rheology of liquid elemental sulfur, the viscosity

standards were fully drained out of the pressure cell and the entire set-up. Pure toluene was loaded

into the system overnight to dissolve possible remnants of the standards before being emptied and

disposed-off into the appropriate organic waste station. This was repeated three times to ensure

that no impurities were left in the system.

200 g of high purity sulfur prill (99.992%) was placed in oven overnight at 135 °C. Before the

MCR 302 was powered on, it was ensured that air valve connected to run the air bearing of the

rheometer was opened and at the required pressure of 6 bar. The set points of the temperature

controllers were set to ca. 140 °C. After the molten sulfur reservoir, HP tubing, fittings and valves

had reached the set temperature; about 80 mL of molten elemental sulfur was carefully poured into

the reservoir from its top while all the valves were closed. With all other valves closed, the plug

on the pressure cell was unscrewed, then V3 (the sulfur reservoir valve, Figure 10) was opened to

allow molten sulfur flow by gravity into the pressure cell until some liquid sulfur had escaped from

the top of the pressure cell. The molten sulfur was allowed to remain in the system overnight. No

data were recorded during this time in other to further clean the system. To empty the sample from

the pressure cell, V3 was closed, then, V4 (the vent valve) was opened and the plug on the pressure

cell was removed to allow the sample to flow out under gravity into waste container. This cleaning

procedure was repeated three times before finally retaining the molten sulfur sample for viscosity

measurements.

65
To finally measure and record the data from this study, fresh molten sulfur was loaded into the

apparatus as described above, the plug on the pressure cell was returned and the gland nut was

tightened. The desired temperature and shear rate were programmed using the Rheoplus software

interface, then the instrument was prompted to start analysis. Each time a fresh molten sulfur

sample was loaded into the pressure cell, the temperature of the sample was held for an

equilibration time of two hours at 5 °C intervals from 120 °C to 280 °C while step-wisely heating

and then, step-wisely cooling the sample. Measurements were made after equilibration at each

temperature interval. For this study, while several viscosity measurements were being made at

constant temperatures, the shear rates were varied. The instrument was programmed to

continuously increase the shear rate on the molten sulfur sample in the pressure cell by increasing

the speed of rotation of the magnetic cylinder which makes the instrument also increase the torque

(M) necessary to sustain this speed. The instrument has an operational M range from 1 nNm to 2

x 108 nNm. The instrument had also been programmed from factory to automatically stop the

increment of γ̇ once the maximum M is closely approached to protect the instrument from being

overloaded.

The entire procedure of equilibrating a fresh liquid sulfur sample at different temperatures while

heating and cooling, as well as making measurements on this same sample before finally draining

out for FTIR spectroscopy analysis is referred to as one charge of experiment.

66
3.7 FT-IR Analysis

Both the fresh sulfur loaded into the apparatus and the analysed molten sulfur drained from the

pressure cell were analysed for impurities using Nicolet iS10 FT-IR spectrometer. This instrument

was calibrated up to 200 ppm by measuring the absorbance of several standard solutions of H2S

dissolved in pure molten sulfur, and then purging before quantifying the total H2S evolved from

the same solution using an ion chromatography technique.24 The FT-IR instrument contains a

thermostated cell maintained at T = 135 °C to keep elemental sulfur sample molten. The cell

window is made of KCl and it has a path length of 5 cm. This FT-IR cell is manufactured at Alberta

Sulfur Research Limited and is widely used in industry.24

67
Chapter Four: RESULTS AND DISUSSION

4.1 Measured Newtonian Viscosity Data of Liquid Elemental Sulfur

The measured Newtonian viscosities were averaged across a shear rate range of 1 to 3500 s-1 at

each temperature. The data obtained for the first four charges of sulfur into the apparatus are given

in the Tables 6, 7, 8 and 9 respectively. The stated uncertainties in η, δ, are estimated using the

standard deviation and a 95.5 % confidence interval, while the uncertainties in T were estimated

to be ± 0.1 °C at 95.5 % confidence interval for the entire measurements.

Table 6. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 1st charge of sulfur into the apparatus.

T / °C η / cP δ / cP

165.0 5600.4 86.3


170.0 17470.7 253.8
175.0 24794.4 88.2
180.0 30134.5 156.7
185.0 32049.4 82.0
190.0 30138.8 76.8
195.0 25157.8 57.0
200.0 20106.2 24.5
205.0 16088.0 21.5
210.0 11975.5 28.3
210.0 8670.4 40.7
205.0 7131.7 57.2
200.0 5620.2 50.1

68
Table 6. Continued

T / °C η / cP δ / cP

195.0 4105.5 66.9


190.0 2647.3 51.9
185.0 1463.8 63.6
180.0 659.8 50.4
175.0 198.3 1.2
170.0 101.7 1.5
165.0 21.8 0.7

Table 7. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 2nd charge of sulfur into the apparatus.

T / °C η / cP δ / cP

150.0 4.7 0.7


155.0 4.0 0.8
160.0 11.7 0.2
165.0 6392.4 186.7
170.0 24176.6 305.7
175.0 40290.9 446.6
180.0 50579.6 223.5
185.0 55782.3 242.6
190.0 56363.5 207.0
195.0 53747.5 191.3
200.0 49121.7 135.9
205.0 43341.6 121.3
210.0 37360.9 70.5

69
Table 7. Continued.

T / °C η / cP δ / cP

215.0 31299.2 73.5


220.0 25824.3 394.7
220.0 20521.0 65.7
215.0 20733.3 62.7
210.0 20582.9 59.1
205.0 19597.5 46.6
200.0 18508.9 103.3
195.0 14854.0 117.9
190.0 10766.9 93.0
185.0 6772.9 74.1
180.0 3379.6 34.5
175.0 1221.3 19.2
170.0 289.1 4.7
165.0 45.8 0.9
160.0 8.6 0.3
155.0 4.8 0.6
150.0 5.0 0.6

70
Table 8. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 3rd charge of sulfur into the apparatus.

T / °C η / cP δ / cP

120.0 11.7 0.4


125.0 10.3 0.3
130.0 9.0 0.1
135.0 7.9 0.3
140.0 6.8 0.2
145.0 5.8 0.2
150.0 5.0 0.2
155.0 4.3 0.2
160.0 5.1 0.2
165.0 8750.7 154.0
170.0 33130.0 481.8
175.0 56918.3 622.4
180.0 73086.8 688.3
185.0 81452.1 537.9
189.8 82793.3 735.7
195.0 79104.5 382.4
200.0 72593.8 538.2
205.0 65071.8 558.8
210.0 55928.5 636.7
215.0 46370.9 672.0
220.0 38693.4 100.3
225.0 32255.6 78.7
230.0 26428.6 105.0
235.0 21366.7 81.5
240.0 13009.4 89.5

71
Table 8. Continued.

T / °C η / cP δ / cP

245.0 13687.3 88.6


240.0 13009.4 89.5
235.0 12796.3 31.3
230.0 13164.3 76.4
225.0 11518.1 72.7
220.0 11058.3 20.8
215.0 9548.4 15.9
210.0 8667.3 12.8
205.0 7484.5 25.9
200.0 6100.1 35.8
195.0 4552.4 55.5
190.0 3070.0 46.6
185.0 1765.0 23.6
180.0 794.6 21.8
175.0 261.3 11.0
170.0 76.3 3.9
165.0 24.3 1.4
160.0 10.7 0.8
155.0 7.3 0.5
150.0 6.7 0.5
145.0 6.9 0.5
140.0 7.6 0.5
135.0 8.4 0.4
130.0 9.5 0.5
125.0 10.8 0.5
120.0 12.1 0.7

72
Table 9. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 4th charge of sulfur into the apparatus.

T / °C η / cP δ / cP

120.0 11.1 0.7


125.0 9.6 0.5
130.0 8.5 0.7
135.0 7.4 0.5
140.0 6.4 0.6
145.0 5.5 0.5
150.0 4.7 0.8
155.0 4.1 0.5
160.0 33.5 14.1
165.0 16702.7 323.1
170.0 45578.5 465.2
175.0 70435.2 634.0
180.0 85771.8 967.3
185.0 92354.8 793.1
190.0 92142.1 699.9
195.0 87065.1 851.7
200.0 79533.3 473.5
205.0 70865.3 1088.5
210.0 61907.3 513.0
215.0 53083.0 955.7
220.0 45091.8 3082.0
225.0 38754.7 548.1
230.0 32509.3 153.2
235.0 27310.1 189.7
235.0 26110.9 181.0

73
Table 9. Continued.

T / °C η / cP δ / cP

230.0 30019.1 351.3


225.0 34215.0 234.0
220.0 38568.5 230.6
215.0 42990.3 216.8
210.0 47119.1 229.9
205.0 50406.3 270.8
200.0 52455.3 260.0
195.0 52386.1 234.0
190.0 49260.2 283.3
185.0 42638.0 218.6
180.0 32484.6 161.6
175.0 19785.3 275.2
170.0 7101.2 166.6
165.0 826.2 26.6
160.0 10.6 1.0
155.0 4.8 0.8
150.0 5.0 0.7
135.0 7.4 0.7
130.0 8.6 0.4
125.0 9.7 0.5
120.0 11.1 0.5

74
The probability distribution characterized by this measurements’ results and their standard

uncertainties, were assumed to be normal. Therefore, the δ of each set of measurements was

estimated to be twice the standard deviation in agreement with the NIST’s convention for coverage

factor, c = 2 ( i.e. 2 x standard deviation), equivalent to 95 % confidence interval.69 However,

because the number of data points per measurement at each temperature varied from five to thirty,

depending on the shear rate at which molten sulfur becomes non-Newtonian, c was therefore

obtained individually for each measurement based on the t-distribution, which takes the degrees

of freedom, v into consideration.

The maximum viscosity, ηmax, of molten sulfur obtained for the 1st, 2nd and 3rd charges charges of

sulfur into the apparatus were approximately 32,000 cP; 56,000 cP and 83,000 cP respectively.

These values are not consistent and are significantly lower than the ηmax value of 93,000 cP

reported by Bacon & Fanelli.2 Finally, for the 4th charge, the heating run matched the data of Bacon

& Fanelli, while the cooling measurements still showed lower ηmax value than the heating

measurements as presented in Figure 16.

75
100000

4th Charge

80000
Bacon & 3rd Charge
Fanelli
Heating
η / cP

60000

2nd Charge
40000 1st Charge

20000

0
140 160 180 200 220 240 260 280
T/°C
Figure 16. Average measured Newtonian viscosity data of molten sulfur as a function of
temperature for the first four charges. ●, heating run of the 1st charge; ○, cooling run of the 1st
charge; ♦, heating run of the 2nd charge; ◊, cooling run of 2nd charge; ■, heating run of the 3rd
charge; □, cooling run of the 3rd charge; ▲, heating run of the 4th charge; ∆, cooling run of 4th
charge ✲, data of Bacon & Fanelli.

4.1.1 The FT-IR Spectroscopy Results

The FT-IR spectroscopy analysis (Figure 17) for the first, second, third and fourth charges shows

total H2S (H2S + H2Sx) contents of 36.1 ppm, 30.7 ppm, 152.0 ppm and 81.0 ppm respectively.

This H2S is believed to be the product of reactions of • Sn• with the hydrocarbons (HCs) of which

the viscosity and rheology standards are made as illustrated in reaction 1 - 2.13 As a result there

was a larger modification of viscosities for the 1st charge where more HCs were available in the

76
pressure cell than the 2nd charge. For this reason, the 1st charge had a higher dissolved total H2S

compared to the 2nd charge. However for the 3rd charge, the maximum temperature of

measurements was increased from 220 °C of the 2nd charge to 250 °C, to further react away

possible impurities in the pressure cell. This produced 152.0 ppm of total H2S in the sample as

shown in Figure 17. It appeared that much of the available HCs reacted away during the 3rd charge

as evident by the lack of a visual peak in the 2900 cm-1 HC region of the IR spectrum for the 4th

charge, making the heating run of 4th charge match the data of Bacon & Fanelli (Figure 16).

2.0
4th Charge;
81.4 ppm (H2S + H2Sx)

1.5
3rd Charge;
152.0 ppm (H2S + H2Sx) 2nd Charge;
30.7 ppm (H2S + H2Sx)

A. U.
1.0

1st Charge;
36.1 ppm (H2S + H2Sx) 0.5
H2Sx
RH

H2S

0.0
3000 2900 2800 2700 2600 2500 2400 2300 2200 2100 2000
Wavenumber / cm-1

Figure 17. FT-IR spectra of the 1st, 2nd, 3rd and 4th charges showing the total H2S (H2S + H2Sx)
content and peaks.

77
The hysteresis between the heating and cooling curves of each charge and the examination of the

FT-IR analysis results led to the thought that trace amounts HCs, which might be present in the

loaded molten sulfur (99.992%) could react at T > 200 °C to produce H2S that modified the

viscosity of the cooling curves.13 Note that the only measurement of Bacon & Fanelli from which

both the heating and cooling experimental data matched using their purified sulfur were carried

out up to a maximum T of 210 °C.2,6 Another possibility is that the equilibration time of two hours

at each T might not be sufficient to establish a dynamic equilibration between all sulfur species

present. To test the validity of the first hypotheses, before further measurements were made, the

pressure cell was loaded with fresh molten sulfur and its temperature held at 295 °C for four days

to ensure that possible remnant of HCs in the pressure cell were reacted away. Subsequent

measurements which were the 5th, 6th, 7th and 8th charges are plotted in Figure 18, while the data

are presented in Appendix B; Tables 12, 13, 14 and 15.

It must be noted however, that the instrument developed electrical faults and could not hold the T

long enough for equilibration on many occasions during the course of measurements for the 5th to

the 8th charge. During this period, the electrical contact points of the pressure cell resistive heating

were modified, by placing contact points outside the heated zone. Corrosion issues were evident

with the manufacturer using brass and silver within the base of the heating mantle. Sulfur vapour

also corroded the electronics of the instrument above the pressure cell and the instrument was

eventually sent back to the manufacturer for repairs. This accounted for the missing data points for

some cooling measurements in Figure 18. The ηmax recorded for all these measurements were

≈ 104,000 cP, which is 10 % higher than ηmax for the 4th charge and the literature value of

Bacon & Fanelli. This could be as a result of a systematic effect in the instrument as it began to

78
fail. Nevertheless, the heating plots of the 5th, 6th, 7th and 8th charges have a good agreement. Even

though, the heating curves still did match the cooling curves for each of these charges, the ηmax of

the cooling run consistently increased. This is attributed to the pressure cell kept getting cleaner of

HCs with each passing charge evident in the total H2S becoming significantly low, reaching a

concentration of 3.8 ppm for the 7th charge and 3.0 ppm for the 8th charge (Table 10).

100000
7th Charge

8th Charge
80000
Bacon &
Fanelli
60000
η / cP

6th Charge

40000
5th Charge

20000

0
140 160 180 200 220 240 260 280
T/°C

Figure 18. Measurements data of average Newtonian viscosity of molten sulfur as a function of
temperature for different charges. ●, heating run of the 5th charge; ○, cooling run of the 5th charge;
♦, heating run of the 6st charge; ◊, cooling run of 6st charge; ■, heating run of the 7th charge; □,
cooling run of the 7th charge; ▲, heating run of the 8th charge; ∆, cooling run of 8th charge ✲,
data of Bacon & Fanelli.

79
Table 10. Total H2S (H2S+H2Sx) content from FTIR analysis of each charge.

Charge Tmax / °C H2S / ppm H2Sx / ppm Total / ppm

1 210.0 9.2 26.8 36.0


2 220.0 6.0 24.7 30.7
3 240.0 82.6 69.4 152.0
4 240.0 1.9 79.5 81.4
5 220.0 0.2 0.5 0.7
6 280.0 1.7 39.2 40.9
7 280.0 0.1 3.7 3.8
8 280.0 0.1 2.9 3.0
9 280.0 0.2 0.7 0.9
10 280.0 0.2 0.7 0.9

It became important to also look at the equilibration time of the molten sulfur and possibility of

trace impurities still present being responsible for the modification of the viscosity of molten sulfur

at high T. Because the equilibration time was two hours at each T before measurements were made,

further measurements were made with equilibration time of four, and six and eight hours while

cooling the sample to check the possibility of recovery of the viscosity within these time durations.

There was no significant effect of increased equilibration time up to eight hours on the cooling

viscosities of molten sulfur. However, for the same sample, repeating the heating run immediately

leads to lower viscosity measurements values for both the heating and the cooling run. By keeping

the sample at 155 °C for 96 hours or more the heating curve was reproduced while the cooling

curve consistently showed increased maximum viscosity. Because smaller amount of HCs were

available to react to modify the viscosity of sulfur at high T as sequential measurements were made

80
on the same sample, leading to smaller viscosity modifications. Whereas for a freshly loaded

molten sulfur, approximately the same concentration of HCs were available to react to modify the

viscosity of molten sulfur. From this, it can be concluded that the equilibration time was not

responsible for the anisotropy of the cooling curve and the small amount of hydrocarbon impurity

in the initial sulfur could be responsible for the anisotropy. The FT-IR analysis of the 9th, 10th and

subsequent charges showed no significant H2S while the total carbon content analysis on the

sample showed the samples having the same total HCs of 50 ± 10 ppm before and after measuring

the sample in the pressure cell of the MCR 302 rheometer. This anisotropy observed can therefore

be attributed to trace amount of HCs that are present in the molten sulfur reacting with sulfur

diradicals to produce H2S which further react to modify the viscosity of the molten sulfur at

T > 210 °C.10,13

4.1.2 Average Measured Newtonian Viscosity Data of Liquid Elemental Sulfur

After the instrument was repaired, important temperatures were selected and programed into the

instrument for measurements. The following temperatures: 120, 155, 175, 187, 200, 240 and

280 °C were chosen because of the abrupt changes to the viscosity of liquid elemental sulfur

observed at these temperatures during the previous eight charges. This would allow the completion

of a number of charges within a shorter time. Also, a shear rate range of 1 - 1500 s-1 was chosen

instead of the 1 - 3500 s-1 used for the previous measurements. This was based on the fact that no

significant changes in the flow behaviour of molten sulfur were observed between 1500 - 3500 s-
1
from the results of the previous charges.

81
The heating runs of the measurements obtained for the 9th - 14th charges were averaged and

presented Table 11. Also, the heating runs measurements data of these charges are plotted in

Figure 19 just like many previous authors.2,29,49 It appears from this study that this was as a result

of the difficulty they encountered in their attempt to reproduce a similar η measurement while

cooling the sample.

Table 11. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures, averaged for the 9th - 14th charges of sulfur into the apparatus.

T / °C η / cP δ / cP

120.0 11.1 0.7


125.0 9.6 0.5
130.0 8.5 0.7
135.0 7.4 0.5
140.0 6.4 0.6
145.0 5.5 0.5
150.0 4.7 0.8
155.0 2.8 1.2
160.0 2.0 0.2
165.1 5415.4 653.4
170.1 30236.2 816.1
175.1 59445.6 950.8
180.1 80551.3 886.4
185.0 91662.4 698.6
187.0 92689.8 489.8
190.1 92864.8 1631.0

82
Table 11. Continued.

T / °C η / cP δ / cP

195.0 91896.6 671.6


200.0 85089.1 936.5
205.0 76695.4 1289.6
210.0 67942.7 88.3
215.0 59163.6 580.0
220.0 51123.5 676.2
225.0 43546.3 618.8
230.0 36573.3 493.9
235.0 31044.0 155.9
240.0 25802.7 570.6
245.1 21535.7 112.5
250.1 17936.2 282.1
255.0 15092.4 246.8
260.1 12454.7 401.3
265.0 10389.4 99.1
270.0 8600.6 230.1
275.1 6983.9 323.6
280.1 6008.0 330.6

83
100000

80000

60000
η / cP

40000

20000

0
140 160 180 200 220 240 260 280
T/°C

Figure 19. Experimental Newtonian viscosity data and models of molten sulfur as a function of T.
✲, experimental data of Bacon & Fanelli;6 –, fittings of Shuai & Meisen;9 □, average experimental
Newtonian viscosity from this study; –, calibrated reptation model of Cates.55

4.1.2.1 Re-Optimising the Modified Reptation Model

There are discrepancies in the literature values of B in equation 2-24. The original value of B by

Cates was therefore kept, while parameters A, C and λ of equations 2-21 to 2-24 were optimised

by minimising the sum of squares using generalized non-linear algorithm. The experimental data

of Ruiz Garcia et. al., Bacon & Fanelli and those obtained from this study were all fitted, the non-

linear algorithm did not converge well.2,29 However, when all four parameters were optimised, a

good fit was found as

84
k  17960 
 7.31013 exp  , (4-1)
 2
 T 

And was plotted in Figure 19. Also, this equation is more sensitive to B than A, and A than C, while

λ was close to unity for each of the fitting attempts in agreement with equation 2-24. The sum of

squares due to error (SSE) between the experimental data and equation 2-24 is 2.5 x 1011, while

the SSE after optimization (equation 4-1) gives 2.8 x 108.

4.2 Non-Newtonian Viscosity Measurements Results of Liquid Elemental Sulfur

4.2.1 Rheology of Sulfur in Mid-Shear Rate Range

During the injection of molten sulfur into underground reservoirs, different γ̇ regimes will be

encountered. Pumping and pipe flow are two important processes with established γ̇ regimes that

are usually encountered during most sulfur transport operations. The mid-shear rate range which

spans from 1 to 1000 s-1, is encountered during pumping processes and pipe flow operations.71,72

The rheology of molten sulfur between 1 to 100 s-1, and 1 to 1000 s-1 at constant T intervals

bounded by 120 °C to 280 °C are summarized in Figure 20 and 21 respectively. The data plotted

are from the 9th - 14th charges of sulfur into the apparatus. Some of the isotherms are omitted for

the sake of clarity.

4.2.1.1 Rheology of Liquid Elemental Sulfur during Pipe Flow

When a sample in the pressure cell thickens and the M of rotation approaches 2 x 108 nNm, the

instrument continues to make measurements without further increment in γ̇ to maintain the

85
viscosity within a safe limit for the instrument. Figure 20 shows the flow behavior of molten sulfur

in shear rate range up to 100 s-1. The γ̇ at which molten sulfur undergoes shear thickening varied

as a function of temperature as can be seen in Figure 20 by the clustering of the red legends. Many

fluids comprise of molecules with irregular shapes and sizes, as shear rate increases, the relative

motion between fluid molecules increases causing them to align and easily move past one another

while reducing the viscosity of the fluids.33 This is the case for Shear thinning and Bingham fluids.

However, for shear thickening fluids, increasing shear rate causes densely packed fluid molecules

to be forced against one another, thereby filling the void spaces in the fluid and leading to increased

viscosity. In molten sulfur, as the average chain length increased, the molecules become more

closely packed causing more entanglements while making the slithering motion between sulfur’s

molecules slower.53,55,60 For each temperature, a shear rate is reached at which the molecules begin

to jam.55 At this point, further shearing at an increased rate would lead to a large resistance to shear

force as observed in the abrupt rise in η of isotherms represented by the red legends of Figure 20

and 21 before the instrument stops the increment of γ̇ in the shear thickening region to prevent

overloading.

86
100000 200 °C

205 °C
Shear thickening region
80000
175 °C

60000 220 °C
η / cP

230 °C
40000

170 °C
245 °C
20000
255 °C
275 °C
165 °C
0
0 20 40 60 80 100
-1
γ̇ / s

Figure 20. Measurement data of the η of molten sulfur as a function of γ̇ up to 100 s-1 at constant T.
□, Newtonian isotherms up to 1000 s-1; ∆, Newtonian isotherms up to 100 s-1 ; ○, shear thickening
isotherms at γ̇ ˂ 100 s-1. Some isotherms are intentionally omitted for clarity.

4.2.1.2 Rheology of Liquid Elemental Sulfur during Pump Operations

Most sulfur pumps operations are within the upper mid shear range which is up to 1000 s-1. The

viscosity measurements data of molten sulfur in this region are presented as a function of shear

rate in Figure 21.

87
100000

165 °C ˂ T ˂ 240 °C
80000

60000 Shear thickening region


η / cP

40000
T ≤ 165 °C
T ≥ 240 °C

20000 T ≤ 165 °C
T ≥ 275 °C

0
0 100 200 300 400 500 600 700 800 900 1000
γ̇ / s-1

Figure 21. Measurement data of the η of molten sulfur as a function of γ̇ up to 1000 s-1 at
constant T. □, Newtonian isotherms up to 1000 s-1; ∆, Newtonian isotherms up to 100 s-1 ; ○, shear
thickening isotherms at γ̇ ˂ 100 s-1. Some isotherms are intentionally omitted for clarity.

The following observations can be made from both Figures 20 and 21. Isotherms of liquid sulfur

at T ≤ 165 °C and T ≥ 275 °C are Newtonian for the entire mid-shear rate range i.e. up to 1000 s-1.

In addition to the relatively low η, ≈ 10,000 cP in this region, this flow behavior implies that molten

sulfur can be handled without the challenge of shear thickening. Likewise in considering the shear

rate range up to 100 s-1, isotherms T ≤ 165 °C and T ≥ 240 °C are Newtonian, making 240 °C to

275 °C a suitable temperature range for shear operations up to 100 s-1.

88
Unlike the flow behaviours discussed above, liquid sulfur becomes shear thickening at different γ̇

known as the critical rate of shear thickening, γ̇ cr, below 100 s-1 within the temperature range

165 °C ˂ T ˂ 240 °C as seen in the sharp increase of η in Figures 21 and 22. A plot of η and γ̇ cr as

a function of T is presented in Figure 22.

100000 400

350
80000
300

250
60000

γ̇ cr / s-1
η / cP

200

40000
150

100
20000
50

0 0
140 160 180 200 220 240 260 280
T/°C

Figure 22. Newtonian viscosity data and critical rate of shear thickening of molten sulfur as
functions of T. □, average experimental Newtonian viscosity from this study; –, calibrated
reptation model of Cates;55 ○, average critical rate of shear thickening from this study.

89
The γ̇ cr of liquid sulfur decreases from the on-set of its λ-transition at T ≈ 160 °C as η rise and T

increases up to 187 °C. The minimum γ̇ cr for the entire sulfur’s liquid temperature range is 30 s-1,

observed at T = 187 °C and η ≈ 92,300 cP. This coincides with the T at which molten sulfur

possesses the longest average chain lengths.52,73 From this point, the γ̇ cr of liquid sulfur begins to

increase as the T increases and η decreases.

90
Chapter Five: CONCLUSIONS AND FUTURE DIRECTIONS

The difficulty many authors encountered in establishing the viscosity data of liquid elemental

sulfur has been a result of the impurities in the sulfur.5,10,11 Bacon & Fanelli established a rigorous

procedure with which they eliminated most of these impurities before making viscosity

measurements, which even today are highly acceptable in the literature. However, it has been over

seventy years since their data were published. A new Anton-Paar MCR 302 rheometer was

commissioned and set up to validate the viscosity data of molten sulfur and for the first time,

investigate its rheometric properties. Viscosity and rheology standards were measured using this

instrument with certified values found well within the 99.7 % confidence interval for all the

standards except the low viscosity N600, with measurements within 99.5 % confidence interval of

certified values for the viscosity range of interest. The sulfur prill used for this study has a purity

of 99.992 % in order to get reliable results. However, the measured standards were difficult to

clean out of the pressure cell and therefore contaminated it. The hydrocarbons of which these

standards were made reacted to produce H2S which modified the viscosity of the measured molten

sulfur samples until the 4th charge where the pressure cell was clean enough that the measurements

matched viscosity data of Bacon & Fanelli.2 Despite this breakthrough for the 4th charge, the

viscosity measurements data for the cooling run were still lower than those for the heating run,

possibly as a result of equilibration or reactions at high temperature.

This anisotropy persisted even after the next four charges of loading fresh molten sulfur samples

into the system, while the sequential heating measurements still showed a good repeatability. The

equilibration time was varied from two to eight hours while cooling the liquid sulfur samples to

91
check the possibility of recovery of the viscosity within these time durations. There was no

significant effect of increased equilibration time up to eight hours on the cooling viscosities of

molten sulfur. The FT-IR analysis of 9th to the 14th charges showed that no significant H2S was

produced in the sample while the total carbon content analysis on the sample also showed the

samples having the same total carbon content of 50 ± 10 ppm before and after measuring the

samples in the pressure cell of the MCR 302 rheometer. The anisotropy observed in each charge

even after the pressure cell appeared to be clean of impurities can therefore be attributed to trace

amount of HCs in the molten sulfur reacting at T > 210 °C. The Newtonian sulfur viscosity data

reported in this study is the average viscosity measurements for the heating runs of the 9th - 14th

charges. The uncertainty in the viscosities were reported at 95.5% confidence interval.

In this study it was also established that liquid sulfur is not Newtonian over its entire liquid range.

The shear rate at which molten sulfur becomes shear thickening varies from ≈ 30 to 100 s-1 with

the minimum shear rate of ≈ 30 s-1 at 187 ° C. Even though, the shear rates applicable for molten

sulfur injection cannot be fully established in this study, two important processes with established

shear rate ranges, and relevant to most molten sulfur transport processes were examined. These

processes fall within the mid-shear rate range: the liquid pipeline transportation (1 to 100 s-1) and

liquid pumping processes (1 to 1000 s-1). The following conclusions can be drawn from the study

of the rheology of molten sulfur in the mid-shear rate range:

(1) Isotherms at T ≤ 165 °C and T ≥ 275 °C appeared to be Newtonian up to 1000 s-1.

(2) Isotherms at T ≤ 165 °C and T ≥ 240 °C appeared to be Newtonian up to 100 s-1.

92
(3) Isotherms between 165 °C ˂ T ˂ 240 °C appeared to be shear thickening at different γ̇ below

100 s-1.

The immediate follow-up work is to obtain sulfur better than 99.992 % to confirm that trace HCs

were entirely responsible for the anisotropy observed in this study. Note that Bacon & Fanelli

reported a reversible viscosity profile with their purified sulfur at 210 °C.6 With even higher purity,

we may be able to show isotropic profiles at higher temperatures.

Once it can be demonstrated that the low-shear viscosities are not anisotropic due to reaction of

impurities, it will be important to revisit the modifying effect of dissolved H2S and H2Sx.

Fanelli (1946), in his work on modifying the viscosity of sulfur, reported the modifying effect of

the viscosity of liquid sulfur as a function of the head pressure of H2S on the molten sulfur. Rubero

(1964), believed Fanelli’s work under-predicted the modifying effect of H2S on the viscosity of

molten sulfur because some of the H2S was lost during the sealing process of the viscometer used

for Fanelli’s measurements.58 He therefore used a different capillary viscometer in a closed system

to measure the modifying effect of H2S on the viscosity of molten sulfur. As he predicted, there

was an increase in the modifying effect of the viscosity of molten sulfur from the same head

pressure of H2S put on molten sulfur compared to Fanelli’s data. However, none of these authors

looked into how much of H2S was actually dissolved into the molten sulfur causing the

modification. It will be desirable to measure these effects using the established FT-IR spectroscopy

technique to analyze the molten sulfur after it is put under a known head pressure of the impurities

and to correlate the lowering of the viscosity that results with known concentrations of the

H2S.24,59,60 It would be interesting to determine if there is still a shear thickening when the H2S has

93
modified low-shear viscosity, i.e., is the shear thickening limit predictable from the low-shear

viscosity? Finally, the previously discussed theoretical models can be tested for H2S modified

elemental sulfur, which should have shorter chain lengths.

Other impurities that have also been suggested but are yet to be studied for their effects on the

viscosity of molten are SO2 and amines (MEA, DEA and MDEA) , which can be present in a Claus

plant where molten sulfur is produced. In addition, the rheology of the sulfur modified by these

impurities will be useful to optimize many flow processes involving molten sulfur.

94
Figure 23. Block diagram of the the VI used to log on pressure measurements from the Keller
pressure transducer to a computer system.

95
Figure 23. Continued.

96
Figure 23. continued.

97
Table 12. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 5th charge of sulfur into the apparatus.

T / °C η / cP δ / cP

150.0 52.2 2.3


155.0 50.1 1.8
160.0 61.9 1.4
165.0 16972.5 723.2
170.0 49559.4 1237.5
175.0 78084.7 2443.4
180.0 96649.0 2529.6
185.0 104461.1 2376.2
190.0 104699.9 513.0
195.0 99383.3 977.0
200.0 91505.8 1214.9
205.0 81723.3 958.7
210.0 71910.6 941.5
215.0 61958.3 828.0
220.0 53373.2 152.0
220.0 51469.8 905.5
220.0 51540.0 441.9
220.0 51327.1 444.3
215.0 58588.6 389.7
210.0 66329.3 610.5
205.0 73938.3 1517.4
200.0 80142.0 1962.8
195.0 84580.1 2224.8

98
Table 12. Continued.

T / °C η / cP δ / cP

190.0 84566.7 2625.0


185.0 78623.5 1805.6
180.0 64997.1 755.3
175.0 44438.2 682.2
170.0 20470.4 117.7
165.0 3104.1 79.0
160.0 61.4 83.1
155.0 54.5 58.5
150.0 54.5 65.5

Table 13. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 6th charge of sulfur into the apparatus.

T / °C η / cP δ / cP

155.0 3.8 1.1


160.0 2.3 0.9
165.0 4234.3 519.5
170.0 31444.9 1441.2
175.0 66831.3 1254.1
180.0 86898.6 465.1
185.0 99477.0 1546.1
190.0 103096.9 2182.3
194.9 101333.0 952.5
200.0 94049.2 1947.0
205.0 85744.7 530.1

99
Table 13. Continued.

T / °C η / cP δ / cP

210.0 77243.7 548.4


215.0 65489.5 116.9
220.0 56269.8 815.0
225.0 48075.5 130.8
230.0 40566.6 473.0
235.0 33818.9 557.6
240.0 28721.7 186.5
245.0 23289.9 914.3
250.0 19202.3 601.8
255.0 16146.7 582.8
260.0 13486.5 604.6
265.0 11259.5 212.0
270.0 9136.5 506.2
275.0 7718.7 274.4
280.0 6466.1 256.6
280.0 6390.1 10.6
275.0 7348.3 393.8
270.0 8556.1 283.2
265.0 9751.1 474.1
260.0 11450.1 10.6
255.0 13219.6 143.0
250.0 15284.3 331.4
245.0 17440.8 473.9
240.0 19829.5 478.7
235.0 22577.1 506.2
230.0 24967.1 1060.1

100
Table 13. Continued.

T / °C η / cP δ / cP

225.0 27553.9 1237.2


220.0 29203.4 549.6
215.0 30750.0 385.8
210.0 31267.6 636.1
205.0 30964.8 464.4
200.0 28563.6 322.0
195.0 24868.2 1327.0
190.0 19384.8 936.5
185.1 12908.7 978.3
180.0 6546.0 241.9
175.1 2544.5 946.9

Table 14. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 7th charge of sulfur into the apparatus.

T / °C η / cP δ / cP

120.0 7.3 4.3


160.0 2.7 0.4
180.0 85636.5 2404.1
185.0 98942.6 4154.2
190.0 102506.7 2103.4
195.0 100007.4 1524.3
200.0 93351.0 1871.1
240.0 29191.9 835.8
280.0 6886.1 357.1

101
Table 14. Continued.

T / °C η / cP δ / cP

280.0 6950.6 91.7


240.0 23643.6 417.0
200.0 47681.5 671.9
195.0 44675.0 391.0
190.0 38953.5 779.1
185.0 29380.0 1242.3
180.0 18638.0 1760.0
160.0 2.6 1.5
140.0 3.4 6.1

Table 15. Average measured Newtonian viscosity data of liquid elemental sulfur at different
temperatures for the 8th charge of sulfur into the apparatus.

T / °C η / cP δ / cP

120.0 10.2 1.1


160.0 2.6 0.8
175.0 69916.1 1932.5
180.0 90775.9 1013.2
185.1 102723.1 276.6
190.1 105658.5 245.3
195.0 103298.8 3415.8
200.1 93967.8 736.3
240.0 29083.2 574.4
240.0 28263.7 183.1
280.0 6836.5 90.8

102
Table 15. Continued.

T / °C η / cP δ / cP

280.0 6739.6 94.7


240.0 25046.4 452.5
200.0 55844.9 420.6
195.0 53183.7 201.4
190.1 47027.4 95.8
185.1 37424.3 489.2
180.0 24458.5 601.9
175.0 10921.4 156.7
160.0 3.7 476.5
120.0 2.9 1.0

103
References

1. Marriott, R. A.; Wan, H. H., Standard Values of Fugacity for Sulfur Which Are Self-Consistent
with the Low-Pressure Phase Diagram. J. Chem. Therm. 2011, 43 (8), 1224-1228.

2. Bacon, R. F. and Fanelli, R. The Viscosity of Sulfur. J. Am. Chem. Soc. 1943, 65, 639-648.

3. Steudel, R.; Steudel, Y.; Wong, M. W. Speciation and Thermodynamics of Sulfur Vapor. Top.
Curr. Chem. 2003, 230, 117-134.

4. R. Steudel, B. Eckert. Solid Sulfur Allotropes. Top. Curr. Chem. 2003, 230, 81-116.

5. Steudel, R. Elemental Sulfur and Sulfur-Rich Compounds I. Springer Berlin Heidelberg: 2003,
230-3, 81-116.

6. Bacon, R. F. and Fanelli, R. Purification of Sulfur. Ind. Eng. Chem. 1942, 34, 1043-1048.

7. Karchmer, J. H. The Analytical Chemistry of Sulfur and its Compounds. In Chemical Analysis;
Elving, P. J.; Kolthoff, I. M., Eds.; Interscience Publishers: 1970; Vol. 29; 8-13.

8. Weinerth, J. Pumping Molten Sulfur: A Challenge for Pump Design, Sealing Systems and
Material of Construction. Friatec AG-Division Rheinhuette Pumpen, Wiesbaden,
Germany.

9. Shuai, X. S. and Meisen, A. New Correlations Predict Physical-Properties of Elemental Sulfur.


Oil Gas J. 1995, 93 (42), 50-55.

10. Fanelli, R. Modifying the Viscosity of Sulfur. Ind. Eng. Chem. 1946, 38 (1), 39-43.

11. Farr, C. C. and Macleod, D. B., On the Viscosity of Sulphur. Proc. Roy. Soc. Lond. Ser. A.
1920, 97 (682), 80-98.

12. Fanelli, R. Solubility of Hydrogen Sulfide in Sulfur. Ind. Eng. Chem. 1949, 41 (9), 2031-2033.

13. Hyne, J. B. Carsul - A Recovered Sulfur Impurity: Formation, Deposition and Control.
Fundamental Sulphur Research Group; 1994.

14. Kutney, G. Sulfur-History, Technology, Application & Industry, 2nd ed.; ChemTec: Toronto,
2013; pp 50-75, 95-115.

15. Clark, P. D. Alternative Energy-Sulfur and Other Commodities: The Inconvenient Reality. In
Alberta Sulphur Research Ltd. Quarterly Bulletin; No. 175 Vol. LII No.3. October
December 2015, pp 20-40.

104
16. Sulfur. Ullmann's Encyclopedia of Industrial Chemistry [Online]; Wiley-VCH Verlag GmbH
&Co Posted December15,2006; http://www.onlinelibrary.wiley.com/doi/10.1002/1435
6007.a25_507.pub2/full (accessed Nov 5, 2016).

17. Paskall, H. G.; Sames, J. A. Basis of the Claus Process. In Sulfur Recovery, 11th ed.; Sulphur
Experts Western Research: Canada, 2008; pp 1-6.

18. Sulfur Recovery. In GPSA Engineering Data Book, 13th ed.; FPS Version: Tulsa, Oklahoma,
2013; Vol. II, pp 1-6 and 13-16.

19. Davis, P.; Marriott, R. A.; Fitzpatrick, E.; Wan, H.; Bernard, F.; Clark, P. D., Fossil Fuel
Development Requires Long-Term Sulfur Strategies. Oil Gas J. 2008, 106 (29), 45.

20. Terry, J. W. The Alberta Sulfur Industry. J. Can. Petrol. Technol. 1984, 23 (5), 65-69.

21. Mineral Commodity Summaries. In U.S. Geological Survey. https://www.minerals.usgs.gov/


minerals/pubs/commodity/sulfur/index.html (accessed Dec 5, 2016).

22. U.S. Geological Survey, Fertilizers: Sustaining Global Food Supplies. 155-99, F. S., Ed. 2017.

23. Rappold, T. A. and Lackner, K. S. Large Scale Disposal of Waste Sulfur: From sulfide fuels
to sulfate sequestration. Energy. 2010, 35 (3), 1368-1380.

24. Marriott, R. A.; Fitzpatrick, E.; Lesage, I. L. The Solubility of H2S in Liquid Sulfur. Fluid
Phase Equilibria. 2008, 269 (1-2), 69-72.

25. Beres, J.; Chakkalakal, F.; McMechen, W.; Sandberg, C.; Controlling Skin Effect Heat Traced
Liquid Sulfur Pipelines with Fiber Optics. Presented at Industry Applications Society
51st Annual Petroleum and Chemical Industry Technical Conference: San Francisco,
CA, USA, Sept 13-15 2004; pp 325-332.

26. Habgood, H. W. Transportation of Sulphur by Pipeline; Research Council of Alberta.


Presented to the sixth Western Regional Conference of the Chemical Institute of Canada:
Trail, B.C., Canada. 1963, 225.

27. King, G. G.; Lawrence, J. E.; Baron, J. J., World's longest sulfur pipeline operating smoothly
after 3 years. Oil Gas J. 1997, 95 (3), 59-65.

28. Shell Canada, Caroline Gas Complex. www.http://shell.ca/can/en_ca/about-us/projects-and-


sites/caroline-gas-complex.html# (accessed September 14, 2016).

29. Ruiz-Garcia, J.; Anderson, E. M.; Greer, S. C. Shear Viscosity of Liquid Sulfur near the
Polymerization Temperature. J. Phys. Chem. 1989, 93.

30. Meyer, B. Elemental Sulfur. Chemical Reviews. 1976, 76 (3), 367-388.

105
31. Peng, D. Y and Zhao, J. Representation of the Vapour Pressures of Sulfur. J. Chem. Therm.
2001, 33 (9), 1121-1131.

32. Strivens, T. A. and Schoff, C. K. Rheometry. Ullmann's Encyclopedia of Industrial Chemistry


[Online]; John Wiley & Sons, Inc. Posted April 15, 2010; http://www.onlinelibrary.wil
ey.com/doi/10.1002/14356007.b06_279.pub2/full (accessed Sep 5, 2016).

33. Schramm, G. Aspects of Rheometry. In A Practical Approach to Rheology and Rheometry.


2nd ed.; 1998; pp 11.

34. Ancey, C. Introduction to Fluid Rheology. Laboratoire Hydraulique Environnementale (LHE);


Ecublens, CH-1015 Lausanne, Suisse, 2005.

35. E473-16: Standard Terminology Relating to Thermal Analysis and Rheology. ASTM
International: West Conshohocken, PA, 2016.

36. Schoff, C. K.; Kamarchik, P. Jr. Rheology and Rheological Measurements. In Kirk-Othmer
Encyclopedia of Chemical Technology [Online]; John Wiley& Sons, Inc. Posted April
15, 2005; http://www.onlinelibrary.wiley.com/doi/10.1002/0471238961.18080515190
30815.a01.pub2/full (accessed Sep 10, 2016)

37. Barnes, H. A.; Maia, M. J. Rheology. In Encyclopedia of Life Support Systems. Vol. I
Rheometry. pp 4-8 and 43-44.

38. Macosko, W. C. Viscous Liquid. In Rheology, Principle, Measurement, and Applications.


Wiley-VCH, Inc.: U.S.A, 1994; pp 78-79.

39. Strivens, T. A., The Shear Thickening Effect in Concentrated Dispersion Systems. J. Col.
Inter. Sc. 1976, 57 (3).

40. Chhabra, R. P.; Richardson, J. F. Non-Newtonian Flow and Applied Rheology-Engineering


Applications [Online]. 2nd ed.; Elsevier: 2008; pp 21-22.

41. Barnes, H. A.; Hutton, J. F.; Walters, K. An Introduction to Rheology. Elsevier Science
Publishers B.V.: Amsterdam, Holland, 1989; Vol. 3, pp 4-8, 43-44.

42. Giacomin, A. J.; Samurkas, T.; Dealy, J. M. A Novel Sliding Plate Rheometer for Molten
Plastics. Polym. Eng. Sci. 1989, 29 (8), 499–504.

43. Mezger, T. G. Rotational Viscometry, Oscillatory Rheometers and Measuring Systems. In


Applied Rheology. 1st ed.; Anton Paar GmbH: Austria, 2015; p 15.

44. Saad, A.; Joseph, R.; Srinivasa, R. Rheology: Tools and Methods (Aviation Fuels with
Improved Fire Safety). Nat. Acad. Sci. 1997, 6.

106
45. Dealy, J. M.; Petersen, J. F; Tee, T. T. A Concentric-cylinder Rheometer for Polymer Melt.
Rheol. Acta. 1973, 12, 550-558.

46. Eley, R. R. Rheology and Viscometry. In Paint and Coating Testing Manual, 15 ed.; Joseph
V. Koleske, Ed.; ASTM International: U.S.A, 2012; pp 415-419.

47. Rotinjanz, L. Viscosity Modification in Liquid Sulphur. Zeitschrift Fur Physikalische Chemie-
Stochiometrie Und Verwandtschaftslehre. 1908, 62 (5), 609-621.

48. Farr, C. C. and Macleod D. B. Some Physical Properties of Gas-Freed Sulphur. Proc. R. Soc.
Lond. Ser. A. 1928, 118 (780), 534-541.

49. Doi, T. Physico-Chemical Properties of Sulfur. Rev. Phy. Chem. Jap. 1965, 35 (1), 18.

50. Gee, G. The Molecular Complexity of Sulphur in the Liquid and Vapour. Trans. Far. Soc.
1952, 48 (6), 515-526.

51. Powell, R. E. and Eyring, H. The Properties of Liquid Sulfur. J. Am. Chem. Soc. 1943, 65,
648 654.

52. Tobolsky, A. V. and Eisenberg, A. Equilibrium Polymerization of Sulfur. J. Am. Chem. Soc.
1959, 81 (4), 780-782.

53. Eisenberg, A. Viscosity of Liquid Sulfur: A Mechanistic Reinterpretation. Macromolecules.


1969, 2 (1), 44.

54. Rouse, P. E. A Theory of the Linear Viscoelastic Properties of Dilute Solutions of Coiling
Polymers. J. Chem. Phys. 1953, 21 (1272).

55. Cates, M. E. Theory of the Viscosity of Polymeric Liquid Sulfur. Eur. phys. Lett. 1987, 4 (4),
497-502.

56. Ferry, J. D., Dilute Solutions: Molecular Theory and Comparison with Experiments. In
Viscoelastic Properties of Polymers, 3rd ed.; John Wiley & Sons, Inc.: 1980; pp 177-
179, 211.

57. Eisenberg, A.; Teter, L. A., Relaxation Mechanisms in Polymeric Sulfur. J. Phys. Chem. 1967,
71 (7), 2332.

58. Rubero, P. A. Effect of Hydrogen Sulfide on the Viscosity of Sulfur. J. Chem. Eng. Dat. 1964,
9 (4), 481-4.

59. Wiewiorowski, T. K. and Touro, F. J., Sulfur-Hydrogen Sulfide System. J. Phys. Chem. 1966,
70 (1), 234.

107
60. Wiewiorowski, T. K.; Matson, R. F.; Hodges, C. T. Molten Sulfur as a Solvent in Infrared
Spectrometry. Anal. Chem. 1965, 37 (8P1), 1080.

61. Bishop, R. H., Learning with LabView 8. Prentice Hall: 2007; pp 8.

62. Deering, C. E. Design, Construction, and Calibration of a Vibrating Tube Densimeter for
Volumetric Measurements of Acid Gas Fluids. M.Sc. Thesis, University of Calgary,
March 2015.

63. Tilford C. R. Pressure and vacuum Measurements. In Physical Methods of Chemistry, 2nd ed.;
John Wiley & Sons, Inc.: 1992; pp 121-129.

64. Pressurements Limited. Certificate of Deadweight Tester Series No: 5165-88.

65. Anton paar Rheometer Information Page. The Modular Compact Rheometer Series.
http://www.anton-paar.com/?eID=documentsDownload&document=18378&L=6
(accessed 15th September, 2016).

66. N600. In Certificate of Analysis. Cannon Instrument Company: USA, 2014; pp 2.

67. N450000. In Certificate of Analysis. Cannon Instrument Company: USA, 2014; pp 2.

68. NIST, Standard Reference Material 2490; Certificate of Analysis. United States Department
of Commerce: Gaithersburg, MD 20899, 2011; p 10.

69. Taylor, B. N.; Kuyatt, C. E., Guidelines for Evaluating and Expressing the Uncertainty of
NIST Measurement Results. U. S. Department of Commerce Technology
Administration; Ed. Physics Laboratory, NIST: Gaithersburg, MD 20899-0001, 1994.

70. Lesage, K. L. and Davis, P. M. Report on the Analysis of Formed Air Prilled Elemental Sulfur
for Keyera Energy. Alberta Sulfur Research Limited. Calgary, Alberta, 2012.

71. Eley, R. R. Rheology and Viscometry. In Paint and Coating Testing Manual, 15th ed.; Joseph
V. Koleske, Ed. ASTM International: U.S.A, 2012; pp 415-419.

72. Shay, G. D., Thickeners and Rheology Modifiers. In Paint and Coating Testing Manual, 15th
ed.; ASTM International: U.S.A, 2012; pp 341-372.

73. Touro, F. J. and Wiewiorowski, T. K. Viscosity-Chain Length Relationship in Molten Sulfur


Systems. J. Phys. Chem. 1966, 70 (1), 239.

108

Potrebbero piacerti anche