You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/320305626

Hydrate Formation Effects on Slug Flow


Hydrodynamics and Heat Transfer: Wall
Deposition vs. Dispersion Formation

Conference Paper · March 2017

CITATIONS READS

0 56

4 authors:

Carlos Lange Bassani Fausto Arinos de Almeida Barbuto


Federal University of Technology - Paraná/Br… Federal University of Technology - Paraná/Br…
23 PUBLICATIONS 24 CITATIONS 29 PUBLICATIONS 43 CITATIONS

SEE PROFILE SEE PROFILE

Amadeu K. Sum Rigoberto E. M. Morales


Colorado School of Mines Federal University of Technology - Paraná/Br…
231 PUBLICATIONS 4,576 CITATIONS 165 PUBLICATIONS 240 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Transient Heat Transfer Modeling of Gas-Liquid Slug Flow Using Slug Tracking View project

Influence of Gas Hydrates-Like Solid Particles on the Slug Flow Hydrodynamics View project

All content following this page was uploaded by Carlos Lange Bassani on 10 October 2017.

The user has requested enhancement of the downloaded file.


IV Journeys in Multiphase Flows (JEM 2017)
March 27-31, 2017, São Paulo, SP, Brazil
Copyright © 2017 by ABCM
Paper ID: JEM-2017-0015

HYDRATE FORMATION EFFECTS ON SLUG FLOW HYDRODYNAMICS


AND HEAT TRANSFER: WALL DEPOSITION VS. DISPERSION
FORMATION
Carlos L. Bassania, langebassani@gmail.com
Fausto A. A. Barbutoa, fausto_barbuto@yahoo.ca
Amadeu K. Sumb, asum@mines.edu
Rigoberto E. M. Moralesa, rmorales@utfpr.edu.br
a
Multiphase Flow Research Center (NUEM), Federal University of Technology – Paraná (UTFPR), Rua Deputado Heitor Alencar
Furtado 5000, Bloco N, CEP 81280-340, Curitiba/PR, Brazil.
b
Hydrates Energy Innovation Laboratory, Chemical and Biological Engineering Department, Colorado School of Mines, 1500
Illinois St., Golden, CO 8040, USA.

Abstract. Pipe blockage due to gas hydrate formation is a main concern to the oil and gas industry due to the costs of
production interruptions. Hydrate formation scenarios are usually found in offshore production pipelines, where oil
and gas flow along the pipeline as a mixture that may also contain sand and brine. The high pressure conditions and
the heat transfer with the external medium – the ocean – may create the necessary conditions for hydrate formation.
Hydrates may form: (i) as a deposit on the pipe inner wall, where the temperature gradient is higher and the wall
imperfections trigger the nucleation process; or (ii) in the gas-water interface, where a more effective contact between
the phases occurs, forming a hydrate-in-liquid dispersion. This work gathers three years of study from NUEM –
Multiphase Flow Research Center on slug flow modeling using a mechanistic approach and shows the main equations,
results and discussions of: (i) slug flows with heat transfer; (ii) mass transfer and heat generation during hydrate
formation; (iii) hydrate formation as a layer deposited on the pipe wall; and (iv) hydrate formation as a dispersion,
thus creating a new flowing phase. A theoretical discussion about the possible mechanisms of transition between
dispersion and wall deposition is also included in the article, focusing on: (i) whether the hydrate dispersion deposits
and (ii) whether hydrate particles may detach from the wall to suspend in the liquid as a dispersion.

Keywords: flow assurance, hydrates, slug flow, heat and mass transfer.

NOMENCLATURE

Roman letters z Pipe axial coordinate [m]


A Cross sectional area [m2] Z Compressibility factor [-]
c Specific heat [J/(kg.K)]
D Pipe inner diameter [m] Greek letters
d Hydrate particle diameter [m]  Pipe inclination [rad]
dm/dt Mass variation rate [kg/s]  Hydrate layer thickness [m]
freq Slug flow frequency [Hz] H Hydration number [-]
h Heat transfer coefficient [W/(m2.K)]
hH Hydrate enthalpy of formation [J/kg]
 Thermal scooping factor [-]
 Viscosity [Pa.s]
j Phase superficial velocity [m/s]
J Mixture superficial velocity [m/s]  Density [kg/m3]
k Thermal conductivity [W/(m.K)]  Shear stress [Pa]
K Head loss coefficient [-]  Phase (   L; G; H )
L Length [m]  Slug region (  B; S )
m Mass flow rate [kg/s]
M Molar mass [kg/kmol] Indexes
P Pressure [Pa] ( ) L Hydrate in water dispersion
R Phase volumetric fraction [-]
S Wetted perimeter [m] ()* Parameter after the consideration of a hydrate
t Time [s] layer deposited in the pipe inner wall
T Temperature [K] B Bubble region
THeq Hydrate formation equilibrium temperature [K] eq Equilibrium
U Real velocity [m/s] ext External medium
VD Critical deposition velocity [m/s] i Gas-water interface
Vt Terminal velocity [m/s] in Pipe inlet
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

G Gas S Slug region


H Hydrate sub Subcooling
L Liquid T Unit cell translation
m Mixture U Unit cell
P Particle W Wall
ov. Overall

1. INTRODUCTION

Hydrates are crystals formed by gas molecules trapped into cages formed by hydrogen-bonded water molecules
(Sloan and Koh, 2008). Whenever natural gas and water mixtures flow through pipelines at high pressures, hydrates
may form: (i) on the pipe wall, where the temperature gradient is higher and the wall imperfections trigger the
nucleation process; or (ii) on the gas-water interface, where a more effective contact between the phases occurs (Sloan
et al., 2011). In either case, the hydrate formation may change the flow hydrodynamics and heat transfer: (i) by
reducing the flow cross sectional area and creating an insulating hydrate layer, in the first case; or (ii) by precipitating
solid particles and forming a dispersion, in the second case (Zerpa et al., 2013). In both situations, pipe blockage may
occur if hydrate formation is out of control, causing either a partial or a complete pipe obstruction due to: (i) the
reduction of the pipe cross sectional area by the formation of a hydrate layer or (ii) the formation of a massive plug due
to particle agglomeration (Sloan et al., 2011).
Pipe blockage due to hydrate formation is one of the main flow assurance challenges faced by the oil and gas
production companies, especially in offshore production operations in deep and cold waters (Cardoso et al., 2015).
Offshore production is mainly composed by oil, gas, water (brine) and small solid particles such as sand, and the
mixture is often considered to be flowing in the slug flow pattern due to the range of volumetric flow rate of the phases.
Slug flows are characterized by the intermittent passage of elongated bubbles, flowing over a liquid film; and liquid
slug bodies, which may contain dispersed bubbles in its interior (Shoham, 2006). Together, these structures form the so-
called unit cell.
Several approaches have been used to model slug flows, namely: steady-state mechanistic models (Bassani et al.,
2016b; Cook and Behnia, 2000; Shoham, 2006; Taitel and Barnea, 1990), Eulerian transient drift flux models
(Danielson, 2011; Zerpa et al., 2013), Eulerian transient two-fluid models (Issa and Kempf, 2003; Simões et al., 2014),
Lagrangian transient slug tracking models (Medina et al., 2015; Nydal and Banerjee, 1996; Taitel and Barnea, 1998)
and hybrid models (Kjeldby et al., 2013). Fewer studies considering heat transfer exist (Bassani et al., 2016b; Medina et
al., 2015; Simões et al., 2014; Zerpa et al., 2013), but the ones considering mass transfer during hydrate formation are
still in development (Zerpa et al., 2013).
This work gathers studies on slug flow modeling using a steady-state mechanistic approach and shows the main
equations, results and discussions about this topic over the last three years of work in NUEM – Multiphase Flow
Research Center. The purpose is to understand the mathematical modeling and the main phenomena involved in:
(i) slug flow hydrodynamics and heat transfer, (ii) mass transfer and heat generation when hydrates form; (iii) hydrate
formation as a layer deposited on the pipe inner wall; and (iv) hydrate formation as a dispersion, which represents a new
flowing phase. At the end of the article, a theoretical discussion based on literature from other groups is added to
understand the main mechanisms of transition between dispersion and wall deposition, that is: (i) whether the hydrate
dispersion deposits and (ii) whether hydrate particles detach from the wall.

2. SLUG FLOW HYDRODYNAMICS AND HEAT TRANSFER

The combined multiphase flow of the multicomponent mixtures from petroleum offshore production – that is, oil,
gas, water (brine), sand and, circumstantially, hydrates – will be modeled as a combined liquid-gas two-phase slug flow.
Figure 1a presents a depiction of a horizontal slug flow, which is characterized by the intermittent succession of unit
cells constituted of two regions, herein designated by  : the elongated bubble (B) and the slug (S). Each region may
contain both phases – that is, the gas (G) and the liquid (L), designated by  . Each phase inside each region is called a
unit cell structure  , being: the gas in the elongated bubble (GB), the liquid film that flows underneath the elongated
bubble (LB), the liquid in the slug body (LS) and the gas bubbles dispersed in the slug body (GS).
Figure 1b presents the problem characterization for hydrodynamic and heat transfer calculations, as proposed by
Bassani et al. (2016b). The pipeline is divided in nodes spaced by  z . The superficial velocities of the phases ( j ) and
the mixture pressure (P) and temperature (T) are assumed to be known at the pipe inlet. The pipeline is cooled
externally by an infinite medium of nearly constant temperature and heat transfer coefficient – which represents the
ocean at offshore production scenarios. The pipeline is assumed to be horizontal.
Knowing the superficial velocities of the phases, the slug unit cell can be characterized by the Taitel and Barnea’s
(1990) approach. Unit cell characterization is related to the volumetric fraction of the phases in each region of the unit
cell, as well as the region lengths. The unit cell geometry is paramount for predicting the real velocities of the phases in
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

each region of the unit cell by the mass balance. The real velocities are used to find: (i) the shear stresses of the
structures for estimating pressure in the next node by the momentum balance, eq. (1) (Bassani et al., 2016b; Cook and
Behnia, 2000; Taitel and Barnea, 1990); and (ii) the heat transfer coefficient of the structures for estimating the mixture
temperature in the next node by the energy balance, eq. (2) (Bassani et al., 2016b). The main phenomena involved in
each conservation equation are indicated inside the equations.

 
 2
 S L  S   S   S  L U  UT   z
P(n 1)  P n    LS LS S  LB LB GB GB i i B  K  L LB  (1)
A LU A LU 2 LU
       
 slug
friction in the
region
friction in the elongated buble region head loss in the
elongated bubble rear


  inheatthetransfer heat transfer scooping


 
       phenomenon
  
film in the slug

 h S L  h S L   m z cL  

T n1  TW  T n   TW  exp    LB LB B LS LS S
  L RLU AU T LU cL
z
 
(2)
     
   
 
liquid heat capacity

Text ; hext

jL ; jG
T;P
Elongated bubble region Slug region

Gas in the
elongated bubble TSf Unit cell geometry Gas superficial velocity correction

r T T
f r Liquid in
LB LS
TB Liquid film
B S
the slug Mass balance Velocities
Gas bubbles dispersed in the slug
Momentum balance P(n+1)
Unit cell TU ; PU  RLS
RLB
Energy balance T(n+1)

(a) (b)
Figure 1 – (a) Unit cell with its respective regions and flow structures, b) problem characterization for slug flow
hydrodynamics and heat transfer calculations.

being L the region lengths, A the cross sectional area occupied by each phase in each unit cell region,   their
shear stresses, S their wetted perimeters, R their volumetric fractions, U their real velocities and h their heat
transfer coefficients. The indexes (n) and (n+1) refer to the nodes of the pipe. Details on the nomenclature are presented
as a separated section in the beginning of this article. The wall temperature TW relates to the external one by means of
ov .
the overall heat transfer coefficient h in each unit cell structure, which by its turn depends on the internal and
external convection and on the wall conduction (Bassani et al., 2016b):

1
 
 
D  hLB
ov .
LB hLSov .
LS   1 D ln  Dext / D  D 
TW  T    Text  T  ; h   
ov .
 (3)
Dext  h L h L  h 2keq Dext hext 
LB U LS U
     
 internal wall conduction external 
 convection convection 

Yet from Fig. 1b, the superficial velocities of the gas phase can be corrected due to pressure and temperature
variations along the pipeline. The liquid superficial velocity is considered constant when assuming an incompressible
fluid hypothesis:
IV Journeys in Multiphase Flows (JEM 2017)

Z  n 1 P n  T n1
jG n1  jG n  ; jL n1  jL n  (4)
Z  n  P n1 T n 

being Z the gas compressibility factor. Equations (1) to (4) demonstrate how to estimate pressure, temperature and
superficial velocities node after node for a case of two-phase gas-liquid slug flow. Since these are the required input
parameters, the routine can be repeated until the end of the pipeline is reached, following an upwind-wise logic, as
shown in Fig. 1b.

3. HYDRATE FORMATION

Hydrate formation implies mainly two phenomena: (i) water and gas consumption and (ii) heat generation, since
hydrate formation has an exothermic nature. Turner (2005) proposes that the gas mass consumption rate dmG/dt due to
hydrate formation depends on the gas-water interfacial surface Ai and on the subcooling of the system Tsub , eq. (5).
This latter is defined as the difference between the hydrate formation equilibrium temperature and the mixture
temperature, Tsub  THeq  T (Zerpa et al., 2013). Hydrates are assumed to form when Tsub  Tsub,crit , where Tsub,crit
is the critical subcooling that activates the first particles nucleation, here assumed as Tsub,crit  3.6 K (Matthews et al.,
2000). The water consumption and the hydrate formation rates, eq. (6), are calculated with respect to the molar masses
of the phases M  and the hydration number  H , i.e., an averaged stoichiometry between the amounts of water and gas
to form the hydrate (Sloan and Koh, 2008).

dmG k 
 k1 exp  2  Ai Tsub (5)
dt T 
dmL M L dmG dmH M dmG
 H ;    H  1 H (6)
dt M G dt dt M G dt

where k1 and k2 are experimental constants. The liquid and gas superficial velocities are recalculated, node by node, due
to the mass transfer when hydrates form (Bassani et al., 2016a). The gas superficial velocity must also take the pressure
and temperature variations into account.

Z  n1 P n  T n1 1 dmG z 1 dmL z


jG n1  jG n   ; jL n1  jL n   (7)
Z  n  P n1 T n  G A dt LU  L A dt LU

From the energy balance standpoint, hydrate formation generates heat. The heat generated can be expressed by the
hydrate enthalpy of formation, in [kJ/kg of gas] (Sloan and Koh, 2008), eq. (8). Therefore, the mixture temperature
estimation between two consecutive nodes should be calculated as shown in eqs. (9) and (10) (Bassani et al., 2016a).

dm
Q H   G hH (8)
dt
p  p  n 
T n1    T n    exp   z  (9)
n  n  m 
 k 
m   L cL ARLU U T LU ; n  q  r ; p  qTW  rTHeq ; q  hLB S LB LB  hLS S LS LS  m z cL ; r  hH k1 Ai exp  2  (10)
 T n1 
 

Figure 2 shows the distribution of temperature (a) and of the mixture superficial velocity (b), evaluated for two
distinct cases: (i) when hydrates do not form (blue line) and (ii) when hydrates form (red line). The model evaluation is
made for a methane-water mixture flowing along a 1.5-km length, 26-mm ID pipeline. The input data for the model
evaluation is specified in Table 1. The following works were used to evaluate the properties of the phases: Haar et al.
(1984) for water, Setzmann and Wagner (1991) for methane and Jung et al. (2010) for hydrates.
The temperature distribution (Fig. 2a) presents a decaying exponential trend, characteristic of the constant external
temperature boundary condition. When hydrate formation is not taken into account (blue line), the mixture temperature
has an asymptote at the external medium temperature (green line). However, the temperature distribution changes when
hydrates are allowed to form. When the decreasing mixture temperature crosses the hydrate equilibrium temperature –
indicated by the red dashed line and evaluated using CSMGem software (Ballard and Sloan, 2004) – the mixture
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

becomes metastable and hydrate formation may occur at any moment. Here, hydrate formation is assumed to form right
after the critical subcooling is reached. This subcooling yields the necessary energy for triggering the nucleation of the
first hydrate particles. Energy is released when hydrate forms,, thus the mixture temperature increases. The mixture
reheats towards the equilibrium temperature, but never to reach it. Actually, a nearly constant subcooling is reached
after the hydrate nucleation. This subcooling is the result of the competition between: (i) the energy released by the
exothermic nature of hydrate formation and (ii) the heat exchange between the mixture and the external medium.
Figure 2b shows the distribution of the mixture superficial velocity. Since the liquid (water) phase is assumed to be
incompressible, variations in the mixture superficial velocity are related to gas expansion/contraction or to the
consumption/formation rates of the phases. When hydrates are not allowed to form (blue line), the gas: (i) contracts due
to temperature drop, thus decreasing the gas (and mixture) superficial velocity; and (ii) expands due to pressure drop,
thus increasing the gas (and mixture) superficial velocity. The first mechanism prevails at the pipeline inlet section,
where the temperature gradient between the mixture and the external medium is the highest. Nearly 300 m downstream,
the mixture superficial velocity presents a minimum, indicating that both mechanisms cancel out themselves. After this
point, the pressure drop mechanism prevails and the mixture velocity increases. When mass transfer due to hydrate
formation is taken into account (red line), the mixture undergoes a slowdown due to gas consumption (a phase with
high specific volume) to form the hydrate (a phase with low specific volume). That is, a mixture volume contraction
takes place, herein related to the mixture slowdown since the system is open (there would be a pressure drop related to
this volume decrease if the system were closed).

Table 1. Input parameters for model evaluation.


Pipe length / ID / width 1.5 km / 26 mm / 1 mm
Pipe inclination Horizontal
Pipe conductivity 30 W/(m·K)
Gas superficial velocity 1 m/s
Liquid superficial velocity 1 m/s
Fluids CH4 / H2O
Pressure at the inlet 10 MPa
Temperature at the inlet 298 K (25oC)
External medium temperature 277 K (4 oC)
External medium heat transfer coefficient 100 W/(m2·K)

(a) 300 (b) 2.2


Mixture superficial velocity [m/s]

Without hydrates
With hydrates 2.1
295
External medium
Temperature [K]

Hydrate equilibrium 2
290
1.9
285
1.8
280 1.7

275 1.6
0 0.3 0.6 0.9 1.2 1.5 0 0.3 0.6 0.9 1.2 1.5
Position [km] Position [km]
Figure 2. Distribution along the pipeline of: (a) mixture temperature and (b) mixture superficial velocity, comparing
cases with and without hydrate formation.

Elongated bubble region Slug region


A B z
Gas in the
elongated bubble

Hydrate
layer

 Liquid film Liquid in the slug Gas bubbles


dispersed in the slug
Hydrates dispersed Hydrates dispersed
A B in the liquid film in the slug

(a) (b)
Figure 3. Two possible cases of hydrate nucleation: (a) at the pipe inner wall as a deposit and (b) at the gas-water
interface, as a homogeneous dispersion.
IV Journeys in Multiphase Flows (JEM 2017)

The aforementioned phenomena involving hydrate formation – that is, mass transfer and heat generation – are
independent on how hydrates form. However, hydrates may form: (i) as a deposit on the pipe wall, where the
temperature gradient is higher and the tiny wall imperfections trigger the nucleation process; or (ii) as a dispersion in
the gas-water interface, where a more effective contact between the phases exists. Figure 3 illustrates both cases. Next,
their effects on slug flow hydrodynamics and heat transfer will be discussed.

3.1 Wall deposition


Hydrates may form on the pipe inner wall, where: (i) the temperature is lower than in the mixture bulk and (ii) the
wall imperfections trigger the nucleation process. Figure 3a shows a hydrate layer  t  deposited symmetrically around
the pipe inner wall, which can be expressed in terms of the hydrate formation rate as:

1 1 dmH
 t   dt (11)
 Dt   H dt

The open flow diameter reduces due to the hydrate wall deposition, eq. (12) (Bassani et al., 2017). The cross
sectional area of the flow is therefore reduced, eq. (13), thus accelerating the phases, eq. (14).

Dt   D0  2 t  (12)
2
 2 t  
A t   1   A0 (13)
 D0
 
2
 2 t  
j t   1   j ,0 (14)
 D0
 

being D0 and A0 the pipe inner diameter and cross sectional areas when no hydrate layer is deposited, with related
superficial velocities of the phases j ,0 . Equations (12) to (14) will affect the pressure gradient, eq. (1), in the following
manners (Bassani et al., 2017):
(i) The friction terms are inversely proportional to the cross sectional area reduction, thus increasing pressure
drop.
(ii) The shear stresses of the unit cell structures are directly proportional to the increase of the superficial
velocities, thus increasing pressure drop.
(iii) The wetted perimeters of the unit cell structures are directly proportional to the diameter reduction, thus
reducing the pressure gradient.
Figure 4 presents the model evaluation for three different constant thicknesses of hydrate layer deposited on the
wall, from 1 to 3 mm. The other input parameters for the model evaluation follow Table 1. The wall deposition is
assumed as already formed – that is, mass transfer and heat generation have already ceased. A case with no hydrate
layer is also plotted for comparison.

(a) 5 (b) 10
No hydrate layer
 = 1 mm
Mixture superficial velocity [m/s]

4  = 2 mm
 = 3 mm 9
Pressure [MPa]

3
8
2

7
1

0 6
0 0.3 0.6 0.9 1.2 1.5 0 0.3 0.6 0.9 1.2 1.5
Position [km] Position [km]
Figure 4 – Distribution along the pipeline of: (a) mixture superficial velocity and (b) pressure, for three different
hydrate wall deposition thicknesses and for a case without the hydrate layer.
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

Figure 4a shows the mixture superficial velocity distribution. After the point where the hydrate layer is present –
that is, after the critical subcooling is reached, near 700 m downstream the pipeline inlet – the mixture undergoes a
sudden acceleration. This acceleration is from ~2 to ~3.5 m/s (~75% augmentation) when a 3 mm hydrate layer is
present (2 D  23%) , and does not take the mixture deceleration due to mass transfer into account, since hydrates
were assumed as already deposited in this simulation. This intensifies the shear stresses, with a consequent pressure
drop increase, as shown in Fig. 4b. The pressure gradient – that is, the slope of Fig. 4b – increases from ~0.9 to
~3.1 kPa/m for the 3 mm layer, which represents ~244% of pressure drop augmentation. This illustrates how sensitive
pressure drop is on the deposited hydrate layer thickness and the high risk related to flow stoppage. The cross sectional
area restriction is probably the hydrate formation phenomenon that influences pressure drops the most.
From the heat transfer standpoint, the hydrate wall deposition represents a conduction thermal resistance (Bassani et
al., 2017). Neglecting contact thermal resistances, the equivalent thermal conductivity between the pipe wall and the
hydrate layer eq. (3)should be used (evaluation of the overall heat transfer coefficient), instead of the wall thermal
conductivity kW :

   
1
D t  ln Dext / Dt   D t  ln D0 / Dt  D ln  Dext / D0  
keq    0 (15)
2  2k H 2kW 
 

On the other hand, the mixture acceleration intensifies the heat exchange between the unit cell structures and the
wall. Since the Reynolds and Nusselt numbers are directly proportional to the velocity of the phases, then an increase in
the mixture heat transfer coefficient with a consequent decrease in the internal convection thermal resistance is
expected. Therefore, the impact on the mixture temperature distribution along the pipeline will be dictated by the
competition between: (i) the increase in the equivalent conductive thermal resistance and (ii) the reduction on the
internal convection thermal resistance.

300
6 000 No hydrate layer
hm [W/(m2.K)]

4 500  = 1 mm
295
3 000  = 2 mm
Temperature [K]

 = 3 mm
1 500 290
(a) Hydrate equilibrium

125
285
hmov. [W/(m2.K)]

100
75
280
50
25 (b) (c)
0 275
0 0.5 1 0.6
1.5 0.9 1.2 0 1.5 0.3
Position [km] Position [km]
Figure 5 – Distribution along the pipeline of: (a) mixture convection heat transfer coefficient, (b) overall heat transfer
coefficient and (c) mixture temperature, for three different hydrate wall deposition thicknesses and the case where no
hydrate layer is deposited.

Figure 5 presents the heat transfer model behavior with hydrate wall deposition, following the input parameters
from Table 1 and the three different hydrate layer thicknesses as presented before. As discussed, the mixture heat
transfer coefficient (Fig. 5a) increases due to the mixture acceleration, from ~3500 to ~6000 W/m2K (~71.4% increase
for   3mm ). However, the equivalent thermal conductivity of the pipe reduces so drastically due to the deposited
insulating layer of hydrate, passing from 30 W/mK (wall conductivity) to ~0.9 W/mK (~97% reduction), that the
overall heat transfer coefficient (Fig. 5b) decreases from ~105 to ~80 W/m2K (~24% reduction). Therefore, the
temperature gradient along the pipeline is reduced when the hydrate layer is present. It is important to notice that a heat
generation case is not shown in Fig. 5, since hydrate formation is assumed to have already ceased.

3.2 Dispersion formation


Hydrates may form in the water-gas interface, such as in the elongated bubble border with the liquid film, or in the
interface between the dispersed bubbles and the liquid slug body. When this occurs, a third phase is introduced in the
flow, as illustrated in Figure 3b, and two new unit cell structures appear: (i) the hydrates dispersed in the slug body and
(ii) the hydrates dispersed in the film. For instance, the hydrate particles are assumed to be homogeneously dispersed
IV Journeys in Multiphase Flows (JEM 2017)

along the unit cell, a valid hypothesis for small particles and low concentrations – that is, at the beginning of hydrate
formation (Joshi, 2012). Thus, the three-phase solid-liquid-gas flow can be treated as a two-phase dispersion-gas flow
(Bassani et al., 2016a). The dispersion properties vary along the pipeline as hydrates forms. The dispersion density is
calculated in terms of the volumetric fraction of water and hydrate, whereas the dispersion viscosity is estimated by
means of the Krieger and Dougherty (1959) correlation:

1.575
R R  R R 
 L   L  L   H H ;  L   L 1  H L  (16)
RL RL  0.63 

being ( ) L the notation used for the dispersion. The properties of the liquid phase in eqs. (1) and (9) should be
substituted by the dispersion one. A new superficial velocity is also introduced in the model – the hydrate one,
estimated via eq. (17) (Bassani et al., 2016a). The dispersion superficial velocity is defined as the sum of the liquid and
hydrate ones, as presented in eq. (18).

1 dmH z
jH  n1  jH  n   (17)
 H A dt LU

jL  jL  jH  j  1 dmL 1 dmH  1 z


L n 1
 jL n      (18)
  L dt  H dt  A LU

Figure 6a presents the dispersion viscosity along the pipelines – evaluated using input parameters of Table 1. When
there is not hydrate formation (dashed line), the liquid viscosity increases along the pipeline due to temperature drop,
from ~0.8 to ~1.5 mPa.s (~87% augmentation) along the 1.5 km-length of the pipeline. When hydrates form, the liquid
viscosification due to the dispersion formation prevails, going from ~1.3 to 2.7 mPa.s (~107% augmentation) from
z ≈ 700 m to the end of the pipeline. However, this viscosification is not sufficient to influence the pressure distribution
along the pipeline, which remains almost the same, as shown in Fig. 6b. The dispersion influence on pressure drop will
be accentuated for higher hydrate volumetric fractions, when the particles distribution may change from homogeneous
to heterogeneous or even forming a moving or stationary bed. However, the model is not yet prepared to handle those
cases, since a homogeneous dispersion hypothesis is used, valid for up to 13% of hydrate volumetric fraction (Joshi,
2012). A theoretical discussion upon the settling of hydrate slurries is made in section 4.

(a) 3 (b)
10
2.5
9.5
Viscosity [mPa.s]

Pressure [MPa]

2
9
1.5 8.5 7.8

7.7
1 8
7.6
0.5 With hydrates 7.5 7.5
Without hydrates 1.4 1.45 1.5
0 7
0 0.3 0.6 0.9 1.2 1.5 0 0.3 0.6 0.9 1.2 1.5
Position [km] Position [km]
Figure 6 – (a) Liquid viscosification due to hydrate dispersion formation and (b) influence on pressure drop when a
homogeneous dispersion is present.

Figure 7a presents the liquid/mixture superficial velocity ratio along the pipeline. Without hydrate formation, jL/J
changes due to gas expansion and contraction. At the inlet section of the pipe (up to z ≈ 300 m), temperature drop is
predominant and the gas contracts, thus jL/J increases. Beyond this point, pressure drop dominates and the gas expands,
thus jL/J decreases. When hydrates form, gas is consumed at a far higher volumetric rate than water, since G   L .
Furthermore,    and then the volume of water consumed nearly equals the volume of hydrate formed, thus jL
H L

remains constant. Therefore, the pseudo-liquid/mixture ratio jL J increases. This increases the slug flow frequency, as
shown in Fig. 7b, since freq  jL J (Gregory and Scott, 1969; Schulkes, 2011). The slug flow frequency – important
in the design of pipelines and phase separators, and especially in predicting corrosion – increases in approximately
16.3% along ~800 m after the hydrates onset point.
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

(a) 0.7 (b) 2.6


With hydrates

Slug flow frequency [Hz]


Without hydrates

0.6 2.4
jL/J [-]

0.5 2.2

0.4 2
0 0.3 0.6 0.9 1.2 1.5 0 0.3 0.6 0.9 1.2 1.5
Position [km] Position [km]
Figure 7 – (a) Liquid/mixture ratio of superficial velocities and (b) slug flow frequency evaluated as proposed by
Schulkes (2011).

4. TRANSITIONS BETWEEN DEPOSITION AND DISPERSION

The technical literature has yet to present a well-established criterion to predict when hydrates may form as a
dispersion or as a wall deposition. Hydrates may also change from dispersion to deposition or vice-versa depending on
the flow conditions. This section brings forward some literature works to theoretically discuss transitions: (i) from
dispersion to deposition, due to decantation of the hydrate particles – also known as settling of slurries; and (ii) from
deposition to dispersion, due to particles detachment from the wall.

4.1 Dispersion-to-deposition transition: settling of slurries


In slurry flows, the critical deposition velocity is defined as the minimum velocity the continuum phase should have
to avoid particle decantation (Peker and Helvaci, 2007). That is, below this velocity, the slurry flow is susceptible to
settling and to forming a stationary bed. The critical deposition velocity depends mainly on the particle diameter, its
drag coefficient and the particle/fluid density ratio. Wilson and Judge (1976) propose:

0.5
 d   
VD  2.0  0.3log  P   2 gD H  1  (19)
 DCD   L 

being VD the critical deposition velocity, dP the particle diameter, D the inner wall diameter, CD the drag coefficient of
the particle and g the gravity acceleration. Equation (19) was modified from the original one by the use of the absolute
value of (  H  L  1) , since the original work focus on particles that are heavier than the carrying fluid, which is not the
case of a hydrate-in-water dispersion. The drag coefficient is dependent on the Reynolds number of the particle, and can
be calculated as proposed by Khan and Richardson (1987) assuming spherical particles and 0.01  ReP  3 105 :

 LVt d P
CD   2.25 ReP0.31  0.36 ReP0.06 
3.45
with ReP  (20)
L

being Vt the terminal velocity of the particle in the continuum fluid, which is expressed in its dimensionless form as
(Zigrang and Sylvester, 1981):

1  2
V*  * 
14.51  1.83( d * )1.5  3.81 (21)
d 
  2
  g  H   
V *  Vt   ; d  dP 
*
 (22)
 g   H   L   2 

If the liquid velocity in one of the regions is below the critical deposition velocity, then the particles are susceptible
to form a deposit:

U LB  VD  hydrate deposition in the film region (23)


U LS  VD  hydrate deposition in the slug region (24)
IV Journeys in Multiphase Flows (JEM 2017)

Figure 8a shows the critical deposition velocity in terms of the hydrate particle diameter. The deposition velocity
increases for bigger particles – that is, the liquid velocity should be higher for the slurry not to settle when its particles
are bigger. Figure 8b shows the mean velocity of the liquid and film regions. When the hydrates nucleate, U LS  2 m s
U  1.7 m s , while the deposition velocity is V  1.8 m s (very small particles). That is, in the film region the
LB D
particles might form a stationary bed, but in the slug region they will most probably flow as a suspension in the liquid
continuous phase.
As the hydrates continue to form and grow (and agglomerate) along the pipeline, the mixture decelerates due to gas
and water consumption to form a solid (as already discussed in Fig. 2b). This flow deceleration also affects the local
film and slug velocities, as shown in Fig. 8b. Therefore, the hydrate slurry may settle due to two mechanisms: (i) the
deposition velocity increases due to the diameter increase of the particles (or of the agglomerated cluster of particles);
and (ii) the carrying fluid velocity ( U LS and U LB ) decrease due to gas and water consumption to form the hydrate. For
example, U ≈ 1.87 m/s at z ≈ 1 km, which implies that only particles smaller than dP ≈ 0.8 mm will flow as a
LS
dispersion in the liquid phase.

(a) 2 (b) 2.2


Critical deposition velocity [m/s]

Without hydrate
2.1 formation
Liquid mean velocity [m/s]
Slug (ULS)
1.95 2
With hydrate
1.9 formation
1.9 1.8
Film (ULB)
1.7
1.85 1.6
1.5
1.8 1.4
0 1 2 3 4 0 0.3 0.6 0.9 1.2 1.5
Particle diameter [mm] Position [km]
Figure 8 – (a) Critical deposition velocity against the particle diameter and (b) distribution along the pipeline of the
mean liquid velocities at the slug and film.

4.2 Deposition-to-dispersion transition: particle detachment from the wall


The hydrates that are deposited on the wall are submitted to forces due to the mixture flow. Whenever those forces
are strong enough to break the adhesion forces between the hydrate and the wall and to lift the particles, the said
particles will flow with the mixture. The evaluation of those forces is, however, extremely dependent on: (i) the shape
of the hydrate layer, (ii) the liquid velocity profile at each region of the unit cell and (iii) the adhesion between the
hydrate and the wall material. Regrettably, these three topics are still open in literature. For instance, a particle with
spherical shape will be assumed to be deposited in the pipe wall, which may be representative of the hydrate nucleation
onset, but is not representative after the hydrate continues to grow, forming a layer as depicted in Fig. 3a.
Figure 9a shows a force balance over a spherical particle deposited on the pipe wall, following the study of Nicholas
et al. (2009). The particle is attached to the wall due to an adhesion force:

Fa ,H /W  Fa ,H /W R* (25)

where the index H/W stands for the hydrate to wall interaction, R* is the harmonic mean radius between the two
surfaces that are in contact ( R*  d P for a sphere in contact with a plane; Yang et al., 2004) and Fa ,H /W is the adhesion
force normalized by R*. Nicholas et al. (2009) suggest that a value of Fa ,H /W  0.002 N m can be used as a conservative
estimation for the normalized adhesion force (valid for interactions between cyclopentane hydrates and carbon steel).
The particle of Fig. 9a is also submitted to lift (FL), gravity (FG) and buoyancy (FB). Forces in the horizontal
direction will not be discussed in this article, since the interest here is to evaluate when the particles will detach and
incorporate into the mixture and assuming that this movement will be entirely vertical. For a spherical particle
submitted to laminar flow – that is, assuming the particle is inside the laminar sublayer – those forces can be estimated
as:

 L dU L
FL  1.615 L d P2 U L (Saffman apud Nicholas et al., 2009) (26)
dP 2
 L dr dP 2
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation


FB  FG    L   H  g d P3 (27)
6

where U L dP 2
and dU L dr d are the liquid velocity and its derivative evaluated at the center of the particle. To
P /2

evaluate these parameters, it is necessary to know the liquid velocity profile U L ( r ) , whose estimation under two-phase
slug flow is not easily obtained from an analytical standpoint. For instance, a Poiseuille flow is assumed in the slug
region, being the maximum velocity at the centerline of the pipeline the double of the mean slug velocity, U L,max  2U LS
(Fox et al., 2011). Thence, the liquid velocity and its derivative can be evaluated as:

2 2
 r  d 
U L r   U L,max 1    U L d 2  2U LS 1  P  (28)
 R  P
 D
dU L r  r dU L r  U d
 2U L ,max    LS P (29)
dr R dr d /2 D
P

Finally, the particle is going to detach and to lift towards the center of the pipeline whenever the following force
criterion is reached (Nicholas et al., 2009):

FL  FB  FG  Fa ,H /W (30)

which is valid for particles attached in the bottom inner wall – that is, buoyancy and gravity have opposite directions.
Figure 9b shows the evaluation of those forces against the particle diameter deposited on the wall. The adhesion
force grows linearly with the particle diameter, whereas lift, buoyancy and gravity forces grow with d P3 . The lift force
(green line) represents the major contribution for the particle detachment, since buoyancy and gravity forces (red line)
almost cancel out themselves (hydrates have a density similar to that of the water, thus resulting in
 L   H  73 kg m3 ) . The purple line represents the sum of lift, buoyancy and gravity forces – that is, the RHS of
eq. (30). For dP ≈ 1 mm, this combined forces equal the adhesion force (blue line). After that point, the particle will
probably detach from the wall.

FL (lift)
6
Criterion for particle
detachment from the wall
Force [N]

FB (buoyancy) FL + FB - FG > Fa,H/W


4

FG (weight) 2

Fa ,H /W (adhesion force) 0
0.5 1 1.5
Particle diameter [mm]
(a) (b)
Figure 9 – (a) Force balance over a spherical hydrate particle that nucleated on the pipe wall (simplified from Nicholas
et al., 2009). (b) Evaluation of adhesion, buoyancy, gravity and lift forces against the particle diameter.

4.3 Final considerations


Conclusions drawn in sections 4.1 and 4.2 may cause confusion, due to the following two statements:
(i) Hydrate particles with dP > 0.8 mm flowing in the slug body are most likely to settle.
(ii) Hydrate particles with dP > 1 mm are most likely to detach from the wall.
It is important to notice that statement (ii) is based on the fact that the lift force over a particle surpasses the
adhesion force between the hydrate and the wall. However, this lift force is not necessarily sufficient to push the
IV Journeys in Multiphase Flows (JEM 2017)

particle towards the centerline of the flow (since this was not analyzed in the present work). Further analysis to
understand this situation should be carried out in a future work, as well as to verify the validity of the correlations and
assumptions adopted. The analysis made in this work should be faced as an initiative to theoretically understand the
mechanisms of transition between deposition and dispersion of hydrate particles.

5. CONCLUSIONS

The present study used a two-phase gas-liquid slug flow mechanistic model to investigate hydrate formation effects
on hydrodynamics and heat transfer. Hydrate formation is related to mass transfer and heat generation. The gas and
water consumption that lead to the formation of a solid – the hydrate – represents a mixture slowdown. By its turn, heat
generation is competitive to heat transfer with the external medium, and thus the system converges to a nearly constant
subcooling when hydrates form.
Other effects of hydrate formation depend on how it nucleates. Two scenarios exist, wall deposition and dispersion
formation. Wall deposition is related to a cross sectional area restriction, which tends to accelerate the mixture and
increase pressure drop. By its turn, hydrate dispersion formation is related to a third phase in the flow. The three-phase
flow can be assumed as a two-phase flow between the dispersion and the gas if the dispersion is treated as
homogeneous. The dispersion formation is related: (i) to a viscosification of the liquid phase, which however did not
influence pressure drop, probably due to the consideration of a homogeneous dispersion (for up to 13% of hydrate
volumetric fraction); and (ii) to an increase in the pseudo-liquid/mixture superficial velocity ratio, which causes the slug
frequency to increase.
Finally, transitions between dispersion and deposition were theoretically analyzed, being: (i) the settling of slurries,
using the critical deposition velocity in terms of the hydrate particle; and (ii) the particles wall detachment, comparing
the lift, buoyancy and gravity forces with the adhesion force between the hydrate and the wall.

6. ACKNOWLEDGEMENTS

The authors acknowledge the financial support of ANP and FINEP through the Human Resources Program for Oil
and Gas Segment PRH-ANP (PRH 10-UTFPR), TE/CENPES/PETROBRAS (0050.0068718.11.9) and the National
Council for Scientific and Technological Development (CNPq). AKS thanks PETROBRAS for sponsoring his
sabbatical leave at UTFPR during the time part of this study was performed.

7. REFERENCES

Ballard, A.L., Sloan, E.D., 2004. The next generation of hydrate prediction IV: A comparison of available hydrate
prediction programs. Fluid Phase Equilib. 216, 257–270. doi:10.1016/j.fluid.2003.11.004
Bassani, C.L., Barbuto, F.A.A., Sum, A.K., Morales, R.E.M., 2017. Modeling the effects of hydrate wall deposition on
slug flow hydrodynamics and heat transfer. Appl. Therm. Eng. 114, 245–254.
doi:10.1016/j.applthermaleng.2016.11.175
Bassani, C.L., Barbuto, F.A.A., Sum, A.K., Morales, R.E.M., 2016a. A mechanistic approach for horizontal gas-liquid
slug flows with the formation of gas hydrate dispersions, in: 9th International Conference on Multiphase Flow.
Firenze, Italy.
Bassani, C.L., Pereira, F.H.G., Barbuto, F.A.A., Morales, R.E.M., 2016b. Modeling the scooping phenomenon for the
heat transfer in liquid-gas horizontal slug flows. Appl. Therm. Eng. 98, 862–871.
doi:10.1016/j.applthermaleng.2015.12.104
Cardoso, C.A.B.R., Gonçalves, M.A.L., Camargo, R.M.T., 2015. Design options for avoiding hydrates in deep offshore
production. J. Chem. Eng. Data 60, 330–335. doi:dx.doi.org/10.1021/je500601f
Cook, M., Behnia, M., 2000. Pressure drop calculation and modelling of inclined intermittent gas-liquid flow. Chem.
Eng. Sci. 55, 4699–4708. doi:10.1016/S0009-2509(00)00065-8
Danielson, T., 2011. A simple model for hydrodynamic slug flow, in: Offshore Technology Conference. Offshore
Technology Conference, Houston/TX, USA. doi:10.4043/21255-MS
Fox, R.W., Pritchard, P.J., McDonald, A.T., 2011. Fox and McDonald’s Introduction to Fluid Mechanics, 8th ed. Sons,
John Wiley &, Hoboken, USA.
Gregory, G.A., Scott, D.S., 1969. Correlation of liquid slug velocity and frequency in horizontal cocurrent gas-liquid
slug flow. AIChE J. 15, 933–935. doi:10.1002/aic.690150623
Haar, V.L., Gallagher, J.S., Kell, G.S., 1984. NBS/NRC steam tables, 1st ed. Hemisphere Publishing Co.
Issa, R.I., Kempf, M.H.W., 2003. Simulation of slug flow in horizontal and nearly horizontal pipes with the two-fluid
model. Int. J. Multiph. Flow 29, 69–95. doi:10.1016/s0301-9322(02)00127-1
Joshi, S. V, 2012. Experimental investigation and modeling of gas hydrate formation in high water cut producing oil
pipelines. PhD Thesis, Colorado School of Mines, Golden/CO, USA.
Jung, J.W., Espinoza, D.N., Santamarina, J.C., 2010. Properties and phenomena relevant to CH4-CO2 replacement in
C.L. Bassani, F.A.A. Barbuto, A.K. Sum and R.E.M. Morales
Hydrate Formation Effects on Slug Flow Hydrodynamics and Heat Transfer: Wall Deposition vs. Dispersion Formation

hydrate-bearing sediments. J. Geophys. Res. Solid Earth 115, B10102. doi:10.1029/2009JB000812


Khan, A.R., Richardson, J.F., 1987. The Resistance to Motion of a Solid Sphere in a Fluid. Chem. Eng. Commun. 62,
135–150. doi:10.1080/00986448708912056
Kjeldby, T.K., Henkes, R. a W.M., Nydal, O.J., 2013. Lagrangian slug flow modeling and sensitivity on hydrodynamic
slug initiation methods in a severe slugging case. Int. J. Multiph. Flow 53, 29–39.
doi:10.1016/j.ijmultiphaseflow.2013.01.002
Krieger, I.M., Dougherty, T.J., 1959. A mechanism for non-newtonian flow in suspensions of rigid spheres. Trans. Soc.
Rheol. 3, 137–152.
Matthews, P.N., Notz, P.K., Widener, M.W., Prukop, G., 2000. Flow Loop Experiments Determine Hydrate Plugging
Tendencies in the Field. Ann. N. Y. Acad. Sci. 912, 330–338. doi:10.1111/j.1749-6632.2000.tb06787.x
Medina, C.D.P., Bassani, C.L., Cozin, C., Barbuto, F.A.A., Junqueira, S.L.M., Morales, R.E.M., 2015. Numerical
simulation of the heat transfer in fully developed horizontal two-phase slug flows using a slug tracking method. Int.
J. Therm. Sci. 88, 258–266. doi:10.1016/j.ijthermalsci.2014.05.007
Nicholas, J.W., Dieker, L.E., Sloan, E.D., Koh, C.A., 2009. Assessing the feasibility of hydrate deposition on pipeline
walls-Adhesion force measurements of clathrate hydrate particles on carbon steel. J. Colloid Interface Sci. 331, 322–
328. doi:10.1016/j.jcis.2008.11.070
Nydal, O.J., Banerjee, S., 1996. Dynamic slug tracking simulations for gas-liquid flow in pipelines. Chem. Eng.
Commun. 141, 13–39. doi:10.1080/00986449608936408
Peker, S., Helvaci, S., 2007. Solid-liquid two-phase flow, 1st ed. Elsevier Science, Amsterdam, Netherlands.
Schulkes, R., 2011. Slug frequencies revisited, in: 15th International Conference on Multiphase Production Technology.
BHR Group, Cannes, France, pp. 311–325.
Setzmann, U., Wagner, W., 1991. A New Equation of State and Tables of Thermodynamic Properties for Methane
Covering the Range from the Melting Line to 625 K at Pressures up to 1000 MPa. J. Phys. Chem. Ref. Data 20,
1061–1155. doi:10.1063/1.555898
Shoham, O., 2006. Mechanistic modeling of gas-liquid two-phase flow in pipes, 1st ed. Society of Petroleum Engineers,
Richardson, USA.
Simões, E.F., Carneiro, J.N.E., Nieckele, A.O., 2014. Numerical prediction of non-boiling heat transfer in horizontal
stratified and slug flow by the two-fluid model. Int. J. Heat Fluid Flow 47, 135–145.
doi:10.1016/j.ijheatfluidflow.2014.03.005
Sloan, D., Koh, C., Sum, A.K., 2011. Natural gas hydrates in flow assurance, 1st ed. Elsevier Inc., Burlington, USA.
Sloan, E.D., Koh, C.A., 2008. Clathrate hydrates of natural gases, 3rd ed. Taylor & Francis Group, Boca Raton, USA.
Taitel, Y., Barnea, D., 1998. Effect of gas compressibility on a slug tracking model. Chem. Eng. Sci. 53, 2089–2097.
doi:10.1016/S0009-2509(98)00007-4
Taitel, Y., Barnea, D., 1990. A consistent approach for calculating pressure drop in inclined slug flow. Chem. Eng. Sci.
45, 1199–1206. doi:10.1016/0009-2509(90)87113-7
Turner, D.J., 2005. Clathrate hydrate formation in water-in-oil dispersions. PhD Thesis, Colorado School of Mines,
Golden/CO, USA.
Wilson, K.C., Judge, D.G., 1976. New techniques for the scale-up of pilot-plant results to coal slurry pipelines, in:
International Symposium on Freight Pipelines. University of Pennsylvania, pp. 1–29.
Yang, S.O., Kleehammer, D.M., Huo, Z., Sloan, E.D., Miller, K.T., 2004. Temperature dependence of particle-particle
adherence forces in ice and clathrate hydrates. J. Colloid Interface Sci. 277, 335–341. doi:10.1016/j.jcis.2004.04.049
Zerpa, L.E., Rao, I., Aman, Z.M., Danielson, T.J., Koh, C.A., Sloan, E.D., Sum, A.K., 2013. Multiphase flow modeling
of gas hydrates with a simple hydrodynamic slug flow model. Chem. Eng. Sci. 99, 298–304.
doi:10.1016/j.ces.2013.06.016
Zigrang, D., Sylvester, N., 1981. An explicit equation for particle settling velocities in solid–liquid systems. AIChE J. 6,
1043–1044.

8. RESPONSIBILITY NOTICE

The authors are the only responsible for the printed material included in this paper.

View publication stats

You might also like