Sie sind auf Seite 1von 15

Chemical Engineering Science 126 (2015) 730–744

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Ebullated bed fluid dynamics relevant to industrial hydroprocessing


Dominic Pjontek a, Craig A. McKnight b, Jason Wiens b, Arturo Macchi a,n
a
Centre for Catalysis Research and Innovation, Department of Chemical and Biological Engineering, University of Ottawa, 161 Louis Pasteur, Ottawa,
ON, Canada, K1N 6N5
b
Syncrude Canada Ltd., 9421-17 Avenue, Edmonton, AB, Canada, T6N 1H4

H I G H L I G H T S

 Overall phase holdups in an ebullated bed were measured at high gas holdups.
 Scale-down was based on geometric features, flow regimes and dimensionless groups.
 Increased liquid viscosity led to gas recirculation and solid holdups reduction.
 Freeboard gas holdups from solids-free estimate depend on bubble–particle interaction.
 Phase holdup correlations were provided for significant bubble coalescence inhibition.

art ic l e i nf o a b s t r a c t

Article history: This study investigates the overall fluid dynamics of an ebullated bed operating at high gas holdup
Received 30 June 2014 conditions to provide relevant observations for industrial residue hydroprocessors. Scaling approaches
Received in revised form for three-phase fluidized beds were compared specifically for the scale-down of the industrially
25 November 2014
observed high gas holdup conditions. Five dimensionless groups, a binary approach for bubble
Accepted 6 January 2015
coalescence behaviour in multi-component liquids, and geometric considerations are proposed to
Available online 13 January 2015
achieve dynamic similitude. Experiments were carried out in a 101.6 mm diameter co-current gas–
Keywords: liquid–solid fluidized bed operating at 0.1 and 6.5 MPa with liquids that do (e.g., 0.5 wt% aqueous
Fluidization ethanol) and do not (e.g., tap water) significantly inhibit bubble coalescence. A comparison of the overall
Ebullated bed
phase holdups for two sizes of cylindrical particles (dSV of 1.6 and 3.9 mm) at matching dimensionless
Gas holdup
groups provided a preliminary verification of the proposed scaling approach when bubble coalescence
Dynamic similitude
Liquid viscosity was sufficiently and consistently inhibited. The impacts of increased liquid viscosity (e.g., greater vacuum
Hydroprocessing distillation tower residue feed fraction), varying superficial gas velocity (e.g., inlet gas flow rate and gas
entrainment in the liquid recycle line), and varying superficial liquid velocity (e.g., liquid recycle pump
speed) were experimentally studied due to their relevance for industrial ebullated bed hydroprocessors.
For the studied operating conditions, gas holdups in the ebullated bed were much greater when bubble
coalescence was sufficiently inhibited (up to approx. 45 vol%) compared to a coalescing system at
equivalent conditions (up to approx. 25 and 35 vol% for 0.1 and 6.5 MPa, respectively). When increasing
the liquid viscosity of the 0.5 wt% aqueous ethanol, a fraction of the gas was entrained in the liquid
recirculation, increasing gas holdups and exhibiting operational similarities to industrial hydroproces-
sors. After adjusting the gas and liquid flow rates based on the gas recirculation, the increased liquid
viscosity mainly reduced the solid holdups while gas holdups were approximately unchanged. Enhanced
bubble break-up or bubble coalescence due to particles resulted in an overestimation or under-
estimation, respectively, of the freeboard gas holdups from a solids-free estimate based on the bed
region phase holdups, where this model could not capture the complex bubble–particle interactions.
Experimental results at high gas holdups conditions were used to correlate the bed and freeboard phase
holdups based on the proposed dimensionless groups.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction

The performance and optimization of industrial ebullated bed


n
Corresponding author. Tel.: þ 1 613 562 5800x6939; fax: þ 1 613 562 5172. residue hydroprocessors, such as the LC-FinerSM, are highly
E-mail address: arturo.macchi@uottawa.ca (A. Macchi). dependent on the overall fluid dynamic behaviour in the bed

http://dx.doi.org/10.1016/j.ces.2015.01.002
0009-2509/& 2015 Elsevier Ltd. All rights reserved.
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 731

Top Pressure Effluent from the literature (Hughmark, 1967; Tarmy et al., 1984). Safoniuk
Tap
Catalyst Addition Line et al. (1999) hence proposed a scale-down approach based on
Thermowell dynamic similitude using the following dimensionless groups:
Density Detector
Radiation Source 2
gðρL  ρG ÞmL 4 gðρL  ρG ÞdP ρL dP U L
Recycle Pan M¼ ; Eo ¼ ; ReL  S ¼ ; ρP =ρL ; U G =U L
ρL 2 γ G  L 3 γG  L mL
ð1Þ
Normal Bed Density
Level Detectors The method assumed that: (i) gas viscosity was negligible
compared to the liquid viscosity, (ii) equilibrium interfacial proper-
Recycle Line Gas-Liquid ties were sufficient to characterize bubble coalescence behaviour,
Ebullated Bed (iii) gas density was much lower than the liquid and solid
densities,
 therefore it was only included in the buoyancy term
Catalyst
(g ρL  ρG ), and (iv) wall effects could be relaxed above a given
Skin
Thermocouples Withdrawal column-to-particle size (dC =dP ) ratio in the dispersed bubble flow
Grid Plate
regime. When attempting to match the proposed dimensionless
groups for the LC-FinerSM using a cold-flow experimental system
with relaxed geometrical constraints, industrial freeboard gas
Hydrogen and holdups nearly doubled those obtained with the laboratory unit
Recycle Bitumen Feed
Pump (McKnight et al., 2003). The significant discrepancy between
industrial and the cold-flow system was attributed to the follow-
Fig. 1. Schematic of the LC-FinerSM hydroprocessor (modified from McKnight
ing possible reasons:
et al., 2003).

1. internal gas recycle via the liquid return line in the


and freeboard regions. Syncrude’s LC-FinerSM operates at elevated industrial unit,
temperatures and pressures of approximately 440 1C and 11.7 MPa 2. inaccurate measurements of phase physical properties and
(McKnight et al., 2003), respectively, required for residue upgrad- holdups in the industrial unit,
ing. A schematic of the unit is provided in Fig. 1. The inlet gaseous 3. inadequate and/or missing dimensionless groups for the fluid
hydrogen and liquid atmospheric/vacuum distillation tower bot- dynamic scale-down.
toms mixture are heated separately and then fed into the plenum
chamber below the grid (i.e., gas–liquid distributor plate) using a Although the first and second considerations could have sig-
horse-shoe/shroud distributor assembly. The feed is mixed with nificantly influenced the comparison, the large gas holdup differ-
the recycled fluid, mainly consisting of unconverted liquid and ences were also believed to be due to difficulties simulating the
some entrained gas from the freeboard region, before flowing high gas holdup conditions in a cold-flow unit. The authors
through the risers and bubble caps located in the grid plate. Doped suggested based on other experimental studies that the influences
alumina cylindrical catalysts are fluidized by the co-current gas of interfacial phenomena for multi-component liquids (Macchi
and liquid flow, where liquid can be considered the continuous et al., 2001) and increased gas density due to elevated pressures
phase while the hydrogen and catalyst constitute the dispersed (Luo et al., 1999; Macchi et al., 2003; Wilkinson et al., 1992) must
phases. Above the ebullated bed, the liquid is recirculated to the be considered.
plenum chamber using a recycle pan and pump. The liquid Few ebullated bed experimental studies which are relevant to
recirculation provides the necessary flow to fluidize the catalyst the fluid dynamics of the LC-FinerSM can be found in the available
particles while also maintaining temperature uniformity through- literature. Tarmy et al. (1984) and Ishibashi et al. (2001) measured
out the reactor. It should be noted that the catalyst bed level is the gas holdups in pilot scale coal liquefaction slurry bubble
monitored using gamma-ray density detectors (shown in Fig. 1). column reactors (i.e., reduced particle size and liquid flow rates
These measurement devices have also been used to estimate compared to an ebullated bed) operating at pressures up to
freeboard gas holdups based on approximations for the gas and 20 MPa and temperatures up to 450 1C. They observed high gas
liquid densities at the reaction conditions (McKnight et al., 2003). holdups which were attributed to the large kinetic energy of the
Discrepancies between industrial measurements and typical high pressure inlet gas and the presence of surface-active compo-
experimental systems available in the literature generally arise nents. Luo et al. (1999) studied the bubble characteristics in a
from considerable differences in operating conditions, phase slurry bubble column operating at pressures up to 5.62 MPa and
physical properties and column geometries. The high gas holdups using Paratherm NF heat transfer fluid. The authors discussed the
observed in ebullated bed hydroprocessors at industrial operating impact of operating pressure on the bubble break-up behaviour,
conditions (McKnight et al., 2003) are difficult to predict or model thus reducing the maximum stable bubble size. The previous
due to the impacts of operating pressure and various interfacial studies provide useful observations on bubble characteristics at
phenomena. Experimental studies at industrially relevant fluid elevated temperatures and/or pressures, however larger solid
dynamic conditions are thus required to improve their design, particles and increased liquid flow rates must be considered
optimization and regular operation. An appropriate scale-down for an ebullated bed. Fan et al. (1987) and Song et al. (1989)
method for the industrially observed high gas holdups must be investigated a 0.5 wt% aqueous t-pentanol solution in a cold flow
identified, where scaling in general still presents an important ebullated bed at atmospheric pressure. The interfacial phenomena
challenge for gas–liquid–solid fluidized beds. leading to bubble coalescence inhibition increased the gas holdups
McKnight et al. (2003) discussed and identified key objectives to due to the reduced bubble size and rise velocities. Luo et al. (1997)
improve the LC-FinerSM performance, noting that minimizing the studied pressure effects on the hydrodynamics and heat transfer in
bed and freeboard gas holdups requires further investigation to an ebullated bed at pressures up 15.6 MPa using Paratherm NF
maximize pitch conversion. Freeboard gas holdup measurements heat transfer fluid. The results provided valuable information on
in the industrial hydroprocessor (approx. 50 to 60 vol%) were the fluid dynamic behaviour when increasing the pressure, how-
considerably greater than predictions (approx. 15 to 25 vol%) for ever relevant high gas holdups were not observed and the super-
comparable operating conditions based on selected correlations ficial liquid velocity was restricted (UL o0.026 m/s). Ebullated bed
732 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

experiments in a 29.4 mm diameter column using 1.7 mm glass between the dispersed and coalesced bubble flow regimes. The
beads, diesel fuel and nitrogen at pressures up to 15 MPa resulted authors therefore proposed a scale-up procedure based on the
in increased gas holdups and reduced minimum liquid fluidization overall gas holdups in a bubble column by incorporating the
velocities due to the modified bubble behaviour (Ruiz et al., 2005, dispersed-to-coalesced bubble flow gas transition velocity, similar
2004). Unfortunately, the studied gas and liquid superficial velo- to Krishna et al. (1991). Their experimental results showed that gas
cities ranges for these studies (UG and UL o20 mm/s) did not result holdups were not considerably influenced by column geometry if
in the high gas holdups observed in industrial units. Sanchez et al. the following conditions were met: (i) column diameter larger
(2008) investigated the Safoniuk et al. (1999) scaling approach by than 0.15 m, (ii) column height-to-diameter ratio greater than 5,
comparing the high pressure results of Luo et al. (1997) to an and (iii) gas distributor holes diameter larger than 1 to 2 mm.
atmospheric system with matching dimensionless groups. Discre- Wilkinson et al. (1992) proposed empirical correlations based on
pancies in the bed porosities and gas holdups for both systems the following dimensionless groups: the capillary number
were generally less than 13% likely due to the differing pressures (vB μL γ G  L  1 ), the Morton number assuming negligible gas den-
and foaming characteristics, where the gasoil used in the atmo- sity (γ G  L 3 ρL g  1 μL  4 ), and the liquid–gas density ratio
spheric system appeared to froth/foam even at low gas velocities. (ρL ρG  1 ). The average error of the correlations was approximately
The purpose of this work is to expand on the previous fluid 10% with a maximum error of 40%. Fan et al. (1999) proposed an
dynamic studies to obtain experimental data relevant to industrial empirical correlation for bubble columns and slurry bubble col-
ebullated bed hydroprocessors. First, a dimensional analysis, umns based on three dimensionless numbers: the ratio of the
which considers the effects of pressure and presence of surfac- superficial gas velocity over the rise velocity of the maximum
tants, is used to scale-down the relevant high gas holdup condi- stable bubble (U G vB; max  1  U G 4 ρG γ G  L  1 g  1 ), a modified Mor-
tions via dynamic similitude. The impact of a more viscous liquid ton number for the slurry phase, and the gas–liquid density ratio.
on the overall fluid dynamics will also be examined as the The average error for this correlation was 13% with a maximum
hydroprocessing feed composition can be varied (Rana et al., error of 53%. Behkish et al. (2006) developed a correlation for
2007), where the vacuum distillation tower residue fraction may slurry bubble columns that was not based on dimensionless
be increased relative to the atmospheric tower residue. The groups, but still provides information on relevant physical proper-
relation between freeboard and bed region gas holdups is then ties. The correlation considered the liquid density, gas density,
discussed as the latter is difficult to estimate while the unit is solid density, liquid viscosity, gas–liquid surface tension, particle
operational due to uncertainties in catalyst inventory and density diameter, solid concentration in the slurry, superficial gas velocity,
(McKnight et al., 2003). Finally, the proposed dimensionless system pressure, vapour pressure of the liquid, column diameter,
groups are used to correlate the overall phase holdups under the gas sparger type, and weight fraction of the primary liquid in a
high gas holdup conditions, where data from a previous study mixture. Differences between predicted and experimental values
(Pjontek and Macchi, 2014) is also included. were within an average absolute relative error (AARE) of 20%. It
should be noted that these correlations were based on experi-
ments with no liquid flow (i.e., batch liquid operation).
2. Fluid dynamic scaling via dimensional analysis and The previous approaches focused on the overall fluid
similitude dynamics by examining global gas holdups, which inherently
assumes that local characteristics will be similar if the previous
Overall characteristics, such as global phase holdups, can have a can be achieved. However, experiments by Shaikh and
significant impact on an industrial ebullated bed’s performance by Al-Dahhan (2010) demonstrated that systems with similar over-
directly affecting major design parameters (e.g., bed region liquid all gas holdups can still differ in local characteristics such
holdup affects the reactant residence time and thus single-pass as radial profiles, mixing, and bubble properties. This also
conversion). Other relevant characteristics such as the fluidized agrees with the observations of Macchi et al. (2001), where
bed interface stability and particle mixing can depend on the local differences in the pressure power spectra indicated disparities
fluid dynamic behaviour. As such, scaling methodologies must in bubble coalescence behaviour at similar overall gas holdups
attempt to match the overall and local fluid dynamics of the for single and multi-component liquids. When considering the
studied systems. The selected scaling approach in this study is reported relative errors between correlated and experimental
based on the principle of dynamic similitude where geometric results, it is difficult to conclude whether the bubble coales-
features (i.e., geometric similitude), the fluid flow regime (i.e., cence behaviour in multi-component liquids is well predicted
kinematic similarity), and a set of dimensionless groups are by such correlations.
matched. This requires the identification of all physical parameters For co-current gas–liquid–solid fluidized beds, Larachi et al.
that have a significant impact on the fluid dynamics of the studied (2001) correlated the overall phase holdups using a combination
system. Failing to include an important property can lead to of artificial neural networks and dimensional analysis (ANN-DA
inaccurate results, while the inclusion of an insignificant para- approach). The correlations were developed based on a large data
meter may create unnecessary experiments which should even- set (20,500 experimental phase holdup measurements for New-
tually demonstrate that it is negligible. tonian liquids), where the following operating conditions and
Scale-up methods for bubble column and slurry bubble column phase physical properties were considered: superficial liquid
reactors can provide an initial assessment for ebullated beds as velocity, liquid density, liquid viscosity, gas–liquid surface tension,
bubble characteristics impact the overall and local fluid dynamic gas density, gas viscosity, superficial gas velocity, particle volume-
behaviour in these systems. When reviewing previous bubble equivalent diameter, particle sphericity, particle density, gravita-
column scale-up attempts, Shaikh and Al-Dahhan (2013) noted tional acceleration, column diameter, and a coalescence index
that no method has yet been able to completely model the local (foaming or coalescing). Cross-correlation coefficients of the liquid
and global behaviour. Nonetheless, the proposed methodologies gravity force, liquid viscous force, capillary force, as well as gas,
provide an overview of relevant physical characteristics and liquid and solid inertial forces were examined for the gas holdup,
general considerations when attempting to achieve dynamic liquid holdup and bed void fraction. Optimal assortments of
similitude for a gas–liquid–solid fluidized bed. dimensionless groups for the outputs were also provided. Com-
Wilkinson et al. (1992) noted that prior gas holdup predictions pared to selected correlations from the literature, the ANN-DA
in bubble columns struggled by not accounting for the transition approach of Larachi et al. (2001) resulted in reduced AAREs for the
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 733

studied data bank (AAREs of 28%, 8.5% and 6% for the gas holdup, recirculate the liquid following the gas–liquid separation. Experi-
liquid holdup and bed void, respectively). ments at high gas holdups by Fan et al. (1987) with an annular
When scaling multiphase reactors, a suitable fluid dynamic three-phase fluidized bed, where the inner-to-outer column dia-
comparison between laboratory scale and industrial units of meter ratio was approximately 1:3, obtained comparable high gas
interest still appears to be challenging. Although the previous holdups to Tarmy et al. (1984), which investigated coal liquefaction
correlations were based on data from various experimental stu- using a pilot scale system at high temperatures (450 1C) and
dies, direct validations of scaling approaches by comparing indus- pressures (17 MPa), at similar superficial velocities. Based on the
trial and laboratory systems are relatively scarce. Considering the previous observations and the experimental system column dia-
average errors between predictions and experimental results, meter of 0.1016 m, constraints for wall effects and the impact of an
particularly for the gas holdup, it is difficult to conclude whether internal recycle line are relaxed due to the dispersed bubble flow
the previous correlations would result in suitable local and global regime and small bubble diameters obtained at high gas holdups.
fluid dynamic similitude under the high gas holdup conditions of
interest. As discussed by Shaikh and Al-Dahhan (2013), scaling 2.2. Formation of dimensionless groups
approaches for multiphase reactors are currently more of an art
than science. It is nonetheless believed that by considering the The ANN-DA approach proposed by Larachi et al. (2001) can be
important geometric and fluid dynamic characteristics of a studied used to provide initial considerations for relevant physical proper-
system, in this case the LC-FinerSM, scaling between industrial and ties when attempting to scale-down the industrial high gas holdup
laboratory equipment using a dimensionless approach can still conditions. The following variables are thus considered: liquid
provide relevant observations. density (ρL ), liquid viscosity (μL ), gas–liquid surface tension (γ G  L ),
gas density (ρG ), gas viscosity (μG ), particle volume equivalent
2.1. Geometric similitude for high gas holdup conditions diameter (dV ), particle sphericity (φ), particle density (ρS ), super-
ficial gas velocity (U G ), superficial liquid velocity (U L ), gravitational
Constraints towards strict geometrical similitude are discussed acceleration (g) and a coalescence index. Similar to Safoniuk et al.
for the gas–liquid separation at the outlet, gas–liquid distribution (1999), the impact of gas viscosity is assumed negligible as μG {μL .
into the ebullated bed, and wall effects due to column diameter Particle shape is accounted for by using the volume equivalent
and internal liquid recycle line. diameter and particle sphericity. A previous study demonstrated
Gas entrainment in the LC-FinerSM recycle pan must be con- that the overall hydrodynamics of spheres and cylinders with
sidered as it may contribute to the observed high gas holdups in matching volume-to-surface area ratios (i.e., equal Sauter mean
the freeboard region (McKnight et al., 2003). Conversely, the diameters) can be similar under high gas holdup conditions
experimental system has a two stage gas–liquid separation, where (Pjontek and Macchi, 2014). At these conditions, the gas, liquid
experimental tests have demonstrated negligible gas entrainment and solid holdup average absolute deviation (AAD) between both
when using a 0.5 wt% aqueous ethanol solution to operate at high shapes were below 1.1%.
gas holdups (gas–liquid separation difficulties when increasing the The inclusion of the gas–liquid equilibrium surface tension
liquid viscosity are discussed in the experimental results). The when scaling-down an ebullated containing a multi-component
impact of gas entrainment in the industrial unit’s liquid recycle liquid is problematic. The gas–liquid surface tension evidently
can nonetheless be essentially simulated by increasing the experi- impacts bubble characteristics (e.g., maximum stable bubble size
mental gas flow rate. and hence its rise velocity). For single component liquids, the rise
The gas–liquid distribution in the LC-FinerSM must also be velocity of a bubble can be represented by the Fan–Tsuchiya
considered as the high gas holdups may be due to significant fluid equation (Fan and Tsuchiya, 1990; Fan et al., 1999) while its
shearing, resulting in enhanced bubble break-up. Gas distribution maximum stable bubble size can be estimated based on the
in bubble columns has been shown to have an influence when the Davies–Taylor equation (Davies and Taylor, 1950), where the gas–
initial bubble size is small relative to its maximum stable size and liquid surface tension is required. In addition to the previous
when the rate of bubble coalescence is low (Tarmy and relations, bubble dynamics are often quantified using the Morton
Coulaloglou, 1992). The impact of the gas distributor is reduced and Eötvös dimensionless numbers, which are again dependent on
at high rates of bubble coalescence as bubbles can reach their the gas–liquid surface tension.
maximum stable size with a sufficient column aspect ratio (hC =dC ). Conflicting results have however been observed when investi-
In the LC-FinerSM, the feed liquid and gas are delivered in a horse- gating the impact of the gas–liquid surface tension on the overall
shoe/shroud distributor assembly and combined with the recycled fluid dynamics of bubble columns or gas–liquid–solid fluidized
liquid before passing through the risers and bubble caps located in beds that contain multi-component liquids or surfactants. Kelkar
the grid plate (McKnight et al., 2003). In the experimental system, et al. (1983) noted a gas holdup increase in a bubble column for
the gas is therefore injected into the liquid using a porous pipe various dilute aqueous aliphatic alcohol solutions, where the
with openings of 10 mm below the distributor to resemble the reduced equilibrium surface tension due to the added surfactants
shearing experienced by both fluids passing through the perfo- was not sufficient to explain the observed trend. Shah et al. (1985)
rated plate. Moreover, local measurements have shown that observed a significant gas holdup increase in a bubble column for
increasing the liquid flow rate reduced bubble chord lengths and varying water–ethanol concentrations when compared to pure
subsequently lowered bubble rise velocities due to enhanced water (i.e., upper equilibrium surface tension) or pure ethanol (i.e.,
bubble break-up (Pjontek et al., 2014). lower equilibrium surface tension). Wilkinson et al. (1992) cau-
Industrial hydroprocessors have relatively large column dia- tioned the use of equations developed using single-component
meters, thus negating wall effects on the overall phase holdups liquids (i.e., coalescing liquids) for liquid mixtures as this will
when compared to laboratory scale systems. Shah et al. (1982) generally underestimate the overall gas holdups. Gorowara et al.
reported negligible wall effects on the gas holdups in a bubble (1990) presented an approach to estimate the gas holdups in
column when the diameter is larger than 0.10 to 0.15 m, similar to three-phase fluidized beds containing surfactants by grouping
the observations of Wilkinson et al. (1992). Kantarci et al. (2005) liquids into four categories based on the equilibrium and dynamic
suggested that wall effects should generally be considered for surface tensions. This approach was unsuccessful when Dargar and
bubble columns with a diameter below 0.1 m. In addition to the Macchi (2006) observed similar gas holdups in a bubble column
outer wall effects, the LC-FinerSM uses an internal recycle line to and fluidized bed for aqueous solutions containing various
734 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

surface-active compounds. For the previous conditions, the type systems under dynamic similitude. Caution should be exercised
and concentration of surfactant mainly affected the foam stability when studying the local flow behaviour between matching
at the free surface. When discussing physical parameter selection laboratory and industrial scale systems as a reduced column
for the ANN-DA approach, Larachi et al. (2001) noted that the diameter may impact the radial flow characteristics, especially in
impact of coalescence inhibitors on bubble break-up and coales- the heterogeneous/coalesced bubble flow regime. It is nonetheless
cence behaviour in an ebullated bed was not yet well understood. believed that the previous has a less significant impact when
The authors thus used a binary coalescence index to account for investigating the overall phase holdups in the homogeneous/
this phenomenon, where the cross-correlation indicated that dispersed bubble flow regime, which are of interest for this study.
coalescence inhibition resulted in higher overall gas holdups.
Experiments with multi-component liquids in both bubble
columns and three-phase fluidized beds thus suggest that once a 3. Experimental system
liquid mixture sufficiently inhibits bubble coalescence, use of the
equilibrium surface tension appears inappropriate to predict the Experiments were carried out in a gas–liquid–solid fluidization
resulting gas holdups. Since the goal of this study is to scale-down system capable of reaching pressures up to 10 MPa. The column is
the fluid dynamics of an industrial ebullated bed containing a made of stainless steel with an inner diameter of 101.6 mm and a
multi-component liquid, where high gas holdups measured in the maximum expanded bed height of 1.8 m. Three glass viewing
freeboard were indicative of foaming (McKnight et al., 2003), the windows are located above the distributor plate. The gas–liquid
equilibrium gas–liquid surface tension will not be considered for separation occurs via the expanded overflow section at the top of
the dimensionless groups. The addition of a surfactant that yielded the column and then by conveying the liquid into a partitioned
high gas holdup conditions with the studied experimental system liquid storage tank for further degassing prior to being recycled to
(Pjontek and Macchi, 2014) will instead be used to simulate the the bottom of the column. The system was pressurized using
suspected bubble coalescence inhibition, consequently resulting in industrial grade nitrogen cylinders. Global phase holdups were
a binary approach for coalescing or coalescence inhibiting liquids determined using a differential pressure transmitter (Rosemount,
(i.e., providing a lower and upper bound, respectively, for the gas model: 1151DP3S22C6Q4). The reference pressure port for the
holdup behaviour). dynamic pressure drop is located at 95 mm above the distributor
The following variables were therefore selected when scaling plate and subsequent pressure ports are equally spaced by a
gas–liquid–solid ebullated beds at high gas holdups: liquid density distance of 146 mm. A centrifugal pump drives the liquid from
(ρL ), gas density (ρG ), particle density (ρS ), liquid viscosity (μL ), the storage tank to the base of the column and a magnetic flow
gravitational
  acceleration via the particle–liquid buoyancy term meter (Rosemount model: 8732CT12N0) measures the liquid flow
(g ρS  ρL ), average particle size/shape using the Sauter mean rate. Gas was circulated via a single stage reciprocating compres-
diameter (dSV ¼ dV φ), gas superficial velocity (U G ), liquid super- sor, where fluctuations in the gas flow are reduced by gas
ficial velocity (U L ), and a binary index for bubble coalescence dampeners located at the compressor inlet and outlet. A differ-
behaviour (coalescing or coalescence inhibition). The particle ential pressure transducer and orifice plates of varying size,
Sauter mean diameter was chosen as the characteristic length depending on the operating pressure, were used to measure the
and the fundamental dimensions used were mass, length, and gas flow rate. Gas was sparged in the plenum chamber of the
time, resulting in the following dimensionless groups based on the column (i.e., below the distributor plate) via a porous pipe with
Buckingham Pi theorem: openings of 10 μm in diameter. The gas–liquid mixture then
  flowed into the bed through a perforated distributor plate with
ρL dSV U L ρL dV 3 g ρS  ρL ρG ρS U G 23 holes of 3.175 mm diameter. A mesh was used to prevent
ReL  S ¼ ; Ar L  S ¼ ; ; ; ð2Þ
μL μL 2 ρL ρL U L particles from entering the plenum chamber. A schematic of the
experimental system and additional details can be found in
In addition to the dimensionless groups, this approach requires Pjontek and Macchi (2014).
equivalent bubble coalescence behaviour for dynamic similarity (i. Operating conditions and phase physical properties for this
e., coalescing or significantly inhibiting coalescence). The previous study, shown in Table 1, were chosen to provide high gas holdup
dimensionless groups indicate that this scaling approach focuses results relevant to the LC-FinerSM. Uncertainties for the gas and
on matching inertial, viscous, and buoyant forces between both liquid superficial velocities as well as operating pressure were
systems. When examining the resulting dimensionless groups, estimated from measurement fluctuations during experiments.
systems with matching particle–liquid Reynolds (ReL  S ) and Considering the binary approach for coalescing or coalescence
Archimedes (ArL  S ) numbers should exhibit equivalent liquid– inhibiting liquids, tap water or a 0.5 wt% aqueous ethanol solution
solid fluidized bed voidage based on empirical correlations for the were respectively selected. An aqueous ethanol solution was used
terminal particle settling velocity and n index parameter (Khan as it was previously shown to produce an effervescent foam at the
and Richardson, 1989), which are parameters required for the free surface (Dargar and Macchi, 2006), a requirement for the
well-known Richardson and Zaki (1954) correlation. As a result, studied experimental system based on the gas–liquid separation at
the proposed scale-down approach matches the liquid–solid elevated pressures. Uncertainties for the liquid density and visc-
fluidized bed characteristics, while the relevant high gas holdup osity were estimated based on repeated measurements, while
behaviour for this study is accounted for by sufficiently inhibiting experimental temperature variations were also considered for the
bubble coalescence, promoting bubble break-up, and matching the viscosity. Aluminum cylindrical particles were selected to mini-
gas–liquid superficial velocity ratio. mize particle density and size distribution effects, while also
If dynamic similitude between separate systems is achieved, having a length-to-diameter ratio and particle–liquid density ratio
equal dimensionless properties should be obtained. For example, comparable to hydroprocessing catalysts. Particle sizing uncer-
both systems should have matching phase holdups as these tainties were estimated based on measurements for 100 partic-
are already dimensionless parameters. The ratio of the bubble les and particle density uncertainties were based on repeated
diameter-to-characteristic length (i.e., Sauter mean particle dia- measurements. A carboxymethyl cellulose (CMC) sodium salt (low
meter) should be also equal. By multiplying various dimensionless viscosity) was added to increase the liquid viscosity, where a
groups, it can be demonstrated that the bubble Reynolds number 0.8 wt% solution resulted in a viscosity of approximately 4.0 mPa  s
and gas–liquid Archimedes number will also match for both (further discussed in Section 5.1). It should be noted that
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 735

Table 1
Studied operating conditions, phase physical properties and dimensionless groups.

Parameter Symbol Range Units

Superficial liquid velocity UL 75 to 123 ( 7 1%) mm/s


Superficial gas velocity UG 0 to 140 (7 2%) mm/s
Pressure P 0.1 and 6.5 (7  1%) MPa
Column diameter dC 101.6 mm
Temperature T 237 2 1C

Liquid density ρL 998 7 2 kg/m3


Liquid viscosity (H2O) μL (0.957 0.4)  10-3 Pa s
Liquid viscosity (0.8 wt% CMC in H2O) μL (4.07 0.3)  10-3 Pa s
Gas density ρG 1.157 0.03 and 73.7 7 0.7 kg/m3

Particle density ρS 2711 78 kg/m3


Particle diameter dP 3.167 0.03 mm
Particle length LP 7.5 7 0.4 mm
Sphericity φ 0.81 70.05 –
Sauter mean diameter dSV 3.9 70.2 mm

Particle–liquid Reynolds number ReL  S 61 to 450 –


Particle–liquid Archimedes number Ar L  S (0.12 to 2.14) x 106 –
Gas-liquid density ratio ρG/ρL 0.0015 and 0.0740 –
Solid–liquid density ratio ρS =ρL 2.505 and 2.716 –
Gas–liquid superficial velocity ratio U G =U L 0 to 2.0 –

measurements using a Anton Paar Physica MCR 301 Rheometer 5. Experimental results and discussion
indicated that the studied CMC concentrations resulted in New-
tonian rheological behaviour. 5.1. Dynamic similitude test via particle size

The proposed scaling method was used to compare the overall


fluid dynamic behaviour of two systems with differing particles sizes
4. Global phase holdups measurements but matching dimensionless groups. Experimental data from a pre-
vious study (Pjontek and Macchi, 2014) using smaller cylindrical
Global phase holdups were calculated by measuring the aluminum particles (ρS ¼ 264979 kg/m3, dSV ¼1.670.2 mm, and
dynamic pressure drop, where the hydrostatic head of liquid is φ ¼0.8070.08) in water and 0.5 wt% aqueous ethanol were compared
subtracted, throughout the bed and freeboard regions. The bed to larger aluminum cylinders, where modified liquid properties and
height (hB ) was estimated from the intersection of the bed and operating conditions for the latter resulted in matching dimensionless
freeboard dynamic pressure axial profiles, obtained by linear groups. For the larger cylinders, the liquid viscosity was increased by
regression. Visual observations of the bed height were recorded adding 0.8 wt% CMC (μL E0.004 Pa s) to approximately match the
when possible to corroborate the bed height obtained by the particle–liquid Archimedes number of the smaller cylinders.
pressure drop method. Solid holdups (εS ) were calculated knowing
the mass of solids in the fluidized bed (m).

4m
εS ¼ ð3Þ 5.1.1. Liquid–solid fluidized bed
π dC 2 hB ρS Fig. 2 presents the solid holdups in the liquid–solid fluidized
beds as a function of the particle Reynolds number for both sizes.
Neglecting frictional drag on the wall and accelerations of the
Horizontal bars in the previous figure were included to illustrate
phases in the vertical direction, gas holdups in the bed region (εG ) were
the estimated ReL  S uncertainty, where deviations were mainly
measured based on the bed region dynamic pressure axial profile.
due to the particle Sauter mean diameters and liquid viscosities.
  Solid holdups for both particle sizes compared relatively well
ΔP=Δz g  1 þðρS  ρL Þ εS
εG ¼ ð4Þ when considering the ReL  S uncertainty and Ar L  S differences.
ρL  ρG
Experimental data was fitted to a modified-dimensionless form
Liquid holdups in the bed region (εL ) were calculated knowing of the well-known Richardson and Zaki (1954) empirical correla-
that the sum of phase holdups must give unity. Gas holdups in the tion:
freeboard region (εG  FB ) were measured based on the dynamic  
pressure axial profile above the bed. ln ReL  S ¼ n ln ð1  εS Þ þ ln ReL  S;T ð6Þ
 
ΔP=Δz g  1 where the intercept provides the Reynolds number at the terminal
εG  FB ¼ ð5Þ
ρL  ρG settling velocity of a particle accounting for wall effects (ReL  S;T )
and the slope estimates the n index. The least squares fit for the
Phase holdups standard deviations were estimated to provide experimental liquid–solid fluidized bed results gave nE 2.7 and
additional insight on the fluid dynamic behaviour of the bed and ReL  S;T E290. The estimated n index (2.4on o4.7) and Reynolds
freeboard regions. Bars presented in the figures of this study number at the terminal free settling velocity (0.2 oReL  S;T o500)
provide the estimated standard deviations for the overall phase indicated the transition between the Stokes (viscous forces dom-
holdups based on the method presented in Pjontek and Macchi inating) and Newton (inertial forces dominating) settling flow
(2014). regimes.
736 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

5.1.2. Gas–liquid–solid fluidized bed as a function of the gas–liquid superficial velocity ratio. The open
Fig. 3 compares the ebullated bed phase holdups and freeboard data points depict water as the solvent (i.e., coalescing system)
gas holdups for both particle sizes with matching ArL  S and ReL  S while the closed data points represent the 0.5 wt% aqueous
ethanol as the solvent (i.e., coalescence inhibition system). A liquid
superficial velocity of 0.075 m/s for the larger cylinders was
0.6
required to match the particle Reynolds number (ReL  S ¼74) of
dSV (mm) ArL-S
0.55 the smaller aluminum cylinders (UL ¼0.045 m/s). It should be
1.6 1.3 x 105
noted that measurements for the smaller cylinders were not
3.9 1.2 x 105 carried out at elevated pressures due to operating difficulties at
0.5
these conditions, such as significant bed expansions leading to
Solid holdup, εS

0.45 partial blocking of the liquid return line (Pjontek and Macchi,
2014). The comparison was thus carried out at atmospheric
0.4 pressure to match the ρG =ρL ratio.
During the experiments, it was observed that adding CMC to
0.35
water resulted in some interfacial phenomena which inhibited
bubble coalescence at the lower gas velocities. Bubbles readily
0.3
coalesced at the higher gas velocities, although small/micro-
0.25 bubbles were still present and gave the liquid a froth-like appear-
ance. The transition from dispersed to coalesced bubble flow for
0.2 the previous conditions can be readily observed based on the gas
30 50 70 90 110 130 150 holdups in Fig. 3a. The resulting behaviour of the 0.8 wt% aqueous
Reynolds number, Re L-S CMC solution thus prevented a suitable comparison to the smaller
Fig. 2. Solid holdup as a function of particle–liquid Reynolds number for smaller
aluminum cylinders in water since the proposed scaling approach
and larger aluminum cylinders in a liquid–solid fluidized bed with matching is based on a binary consideration for bubble coalescence beha-
dimensionless groups. viour (i.e., comparisons must be made between systems that do or

0.7 ReL-S = 74 0.7 ReL-S = 74

0.6 0.6 coalescence


inhibition / gas recycle
Bed region gas holdup, εG

Freeboard gas holdup, εG-FB

0.5 0.5

coalescence
0.4 0.4 inhibition

0.3 0.3 mixed


behaviour

0.2 0.2

coalescing
0.1 0.1

0 0
0 1 2 3 0 1 2 3
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

0.65 0.65
ReL-S = 74 ReL-S = 74
0.6 coalescence 0.6
Bed region liquid holdup, εL

C CI dSV (mm) ArL-S


0.55 0.55
1.6 1.3 x 105
0.5 3.9 1.2 x 105 0.5
Solid holdup, εS

0.45 0.45

0.4 0.4

0.35 0.35

0.3 0.3

0.25 0.25

0.2 0.2
0 1 2 3 0 1 2 3
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

Fig. 3. Ebullated bed and freeboard phase holdups as a function of gas–liquid superficial velocity ratio for smaller and larger aluminum cylinders in water (i.e., coalescing/
mixed behaviour (C) systems) and 0.5 wt% aqueous ethanol (i.e., coalescence inhibition (CI) systems) at P ¼0.1 MPa.
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 737

0.6 coalescence inhibition 0.6

0.5 0.5

Bed region gas holdup, εG

Freeboard gas holdup, εG-FB


0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

0.45 0.7
ρG / ρL
0.65
0.4 0.002 0.074 ArL-S ReL-S
0.6 21.0 x 105 375

Bed region liquid holdup, εL


1.2 x 105 87
0.35 0.55
Solid holdup, εS

0.5
0.3
0.45

0.25 0.4

0.35
0.2
0.3

0.15 0.25
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

Fig. 4. Ebullated bed phase holdups for the coalescence inhibition systems at varying gas flow rates and liquid viscosity.

do not significantly inhibit bubble coalescence, whereas mixed Overall phase holdups in the ebullated bed were relatively
behaviour is difficult to quantify). comparable for both coalescence inhibition systems in dispersed
Following the addition of CMC to the 0.5 wt% aqueous ethanol, bubble flow. Discrepancies can be observed for the coalescence
it was observed that a fraction of the gas was being recycled with inhibition systems at higher gas flow rates (U G =U L 4 1.55) as the
the liquid recirculation (readily observed after stopping the inlet smaller cylinders transitioned to coalesced bubble flow (Pjontek
gas flow). It is believed that the gas recycle resulted from the and Macchi, 2014). The previous demonstrated the importance of
addition of a second surface-active component and/or the reduced matching the fluid flow regimes (i.e., kinematic similarity) for the
film drainage rate between two adjacent bubbles as it is inversely scaling approach. Bearing in mind the differences in particle size
proportional to the liquid viscosity (Sagert and Quinn, 1978, 1977). (dSV of 1.6 and 3.9 mm), superficial liquid velocities (U L of 45
The resulting foam at the free surface consequently did not and 75 mm/s), and liquid viscosities (μL of 9.5  10-4 and 4.0 
entirely dissipate in the two gas–liquid separation stages (i.e., 10-3 Pa s) for both systems, the similar overall phase holdups
expanded overflow section and partitioned recycle tank) prior to trends in the ebullated bed provided a preliminary confirmation
being recycled to the bottom of the ebullated bed. The gas fraction of the proposed dimensionless scaling approach for similar bubble
in the liquid recycle was experimentally estimated by measuring coalescence behaviours. Freeboard gas holdups observed with the
the dynamic pressure drop in the freeboard region shortly after larger particles were however greater than those obtained with
gas shut-off while maintaining liquid flow. The superficial velocity the smaller cylinders. Improved measurements of the gas fraction
ratios presented in Fig. 3 were adjusted to account for the in the liquid recirculation are required to confirm this trend.
increased gas and reduced liquid superficial velocities due to the
gas recycle, thus providing an estimate of the actual gas and liquid
flow rates in the ebullated bed. At the highest superficial gas
velocity shown in Fig. 3, the fluid recirculation in the experimental 5.2. Effects of gas and liquid superficial velocities and an increased
system had a gas fraction of approximately 17 vol%. The volumetric liquid viscosity
fraction of gas in the liquid recirculation was approximated based
on measurements taken at multiple gas velocities. It is important Measurements including the high gas holdup conditions rele-
to note that the magnetic flow meter on the liquid recycle line vant to industrial ebullated bed hydroprocessors (i.e., coalescence
measures the volumetric fluid flow rate independently of fluid inhibition using the 0.5 wt% aqueous ethanol, before and after
density. CMC addition) were carried out to study the following parameters:
738 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

coalescing / mixed behaviour


0.5 0.5

Bed region gas holdup, εG

Freeboard gas holdup, εG-FB


0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 0.5 1 1.5 0 0.05 0.1 0.15
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

0.45 0.7

0.65
0.4
0.6
Bed region liquid holdup, εL
0.35 0.55
Solid holdup, εS

0.5
0.3
0.45

0.25 0.4
ρG / ρL
0.35 0.002 0.074 ArL-S ReL-S
0.2 21.0 x 105 375
0.3 1.2 x 105 87

0.15 0.25
0 0.5 1 1.5 0 0.5 1 1.5
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

Fig. 5. Ebullated bed phase holdups for the coalescing (water) and mixed behaviour (0.8 wt% aqueous CMC) system at varying gas flow rates and liquid viscosity.

1. increased liquid viscosity due to higher vacuum distillation and water (coalescing/mixed behaviour systems) are presented in
tower residue feed fraction, Figs. 4 and 5, respectively, where the inlet gas flow was varied while
2. varying superficial gas velocity due to the gas feed flow rate or maintaining the liquid flow rate. The effects of bubble coalescence
gas recycle fraction, behaviour and operating pressure with respect to the overall phase
3. varying superficial liquid velocity due to the liquid feed flow holdups have already been discussed for the studied particles in water
rate or liquid recycle. and 0.5 wt% aqueous ethanol in a previous study (Pjontek and
Macchi, 2014). Increasing the pressure allowed the coalescing system
Similar to the results discussed in the previous section, gas to remain in dispersed bubble flow for higher gas superficial velocities
entrainment in the liquid recirculation was observed when adding (refer to Fig. 5a). The addition of CMC to water again resulted in some
CMC to the coalescence inhibition system (0.5 wt% aqueous ethanol coalescence inhibition at low gas velocities whereas bubble coales-
as the solvent). The superficial velocity ratios presented in Figs. 4 and cence occurred at increased gas flow rates, which can be readily
6 were therefore adjusted based on experimental measurements to observed by the greater phase holdup standard deviations (Fig. 5) due
account for the increased gas and reduced liquid superficial velocities to the formation of larger bubbles. Elevated pressures had a less
due to the gas recycle (refer to the method described in Section 5.1.2). significant impact for the aqueous ethanol systems (refer to Fig. 4a),
Although not representative of the high gas holdup behaviour where bubble break-up via the gas–liquid distribution system con-
in industrial ebullated beds, data is also presented for the coales- siderably reduced the average bubble size based on visual observa-
cing and mixed behaviour systems (i.e., water and 0.8 wt% aqueous tions. Photographs at comparable operating conditions in the
CMC, respectively) for comparison purposes. A system that does experimental system can be found elsewhere (Pjontek et al., 2014).
not inhibit bubble coalescence provides a lower bound for the gas Elevated pressure following CMC addition to the 0.5 wt% aqueous
holdups at matching dimensionless groups. Unfortunately, the ethanol was not investigated as the foam layer stability would have
observed surface-active characteristics following CMC addition in likely resulted in liquid entering the gas compressor, where the
water prevented the isolated investigation of increased liquid polymer characteristics of CMC could have damaged the internals.
viscosity in a coalescing system. Fig. 4 demonstrates that the bed and freeboard gas holdups
were not significantly affected by the increased liquid viscosity.
Bed region gas holdups trends indicated operation in the dispersed
5.2.1. Varying inlet gas flow rate bubble flow regime for the 0.5 wt% aqueous ethanol (e.g., rate of
Ebullated bed and freeboard phase holdups before and after CMC increase of εG versus U G =U L and relatively low phase holdup
addition in 0.5 wt% aqueous ethanol (coalescence inhibition systems) standard deviations shown in Fig. 4a). McKnight et al. (2003)
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 739

0.55 0.55
coalescence inhibition
0.5 0.5

Freeboard gas holdup, εG-FB


Bed region gas holdup, εG
0.45 0.45

0.4 0.4

0.35 0.35

0.3 0.3

0.25 0.25

0.2 0.2

0.15 0.15

0.1 0.1
0.6 0.8 1 1.2 0.6 0.8 1 1.2
Superficial velocity ratio, U G / UL Superficial velocity ratio, UG / UL

0.4 0.6

ρG / ρL
320
0.36 0.55 0.002 0.074 ArL-S ReL-S

Bed region liquid holdup, εL


ReL-S 21.0 x 105 320 to 450
1.2 x 105 74 to 104
0.32 0.5
Solid holdup, εS

450

0.28 0.45

74
0.24 0.4
ReL-S

0.2 0.35
104

0.16 0.3
0.6 0.8 1 1.2 0.6 0.8 1 1.2
Superficial velocity ratio, UG / UL Superficial velocity ratio, UG / UL

Fig. 6. Ebullated bed phase holdups for the coalescence inhibition systems at varying liquid flow rates and liquid viscosity.

reported freeboard gas holdups ranging from approximately 50 to 5.2.2. Varying liquid recirculation rate
60 vol% in the industrial hydroprocessor, where similar values Fig. 6 and 7 present the ebullated bed and freeboard phase
were obtained in the coalescence inhibiting system (refer to holdups for the 0.5 wt% aqueous ethanol (coalescence inhibition)
Fig. 4b). Measurements yielded a gas fraction of approximately and water (coalescing/mixed behaviour) systems, respectively,
20 vol% in the liquid recirculation at the highest gas superficial when varying the liquid recirculation flow rate while maintaining
velocity shown in Fig. 4. Freeboard gas holdups at the previous the inlet gas flow rate. Increasing the liquid superficial velocity
conditions increased by approximately 10 vol% with gas entrain- thus reduced the gas–liquid superficial velocity ratio (U G =U L ) and
ment when operating at equal inlet gas flow rates (i.e., without augmented the particle Reynolds numbers, as illustrated in Fig. 6c.
adjusting for gas entrainment in the liquid recirculation). Gas For the 0.5 wt% aqueous ethanol prior to CMC addition, bed region
fractions in the liquid recirculation increased at higher inlet gas gas holdups decreased at higher liquid superficial velocities
flow rates, where a similar trend in the industrial unit has been (Fig. 6a), where a reduction of 3.7 vol% was observed for the
previously observed (McKnight et al., 2003). studied ReL  S range. CMC addition to the coalescence inhibition
Comparing 0.5 wt% aqueous ethanol (Fig. 4d) and water (Fig. 5d), system increased the gas–liquid superficial velocity ratio due to
it is apparent that the high gas holdup conditions resulted in lower gas entrainment (approx. 15 vol%) in the liquid recirculation (i.e.,
overall bed region liquid phase holdup, therefore reducing the liquid higher gas and reduced liquid flow rates in the ebullated bed for
residence time. Fig. 4c demonstrates that increasing the liquid equivalent inlet gas and fluid recirculation flow rates), subse-
viscosity mainly reduced the solid holdups (i.e., greater bed expan- quently increasing U G =U L as well as the bed and freeboard gas
sion) due to the relative gain in viscous forces compared to particle holdups. Bed region gas holdups for the coalescing system (Fig. 7a)
inertial (ReL  S reduction) and gravitational (ArL  S reduction) forces. increased for higher liquid flow rates at atmospheric pressure
Experimental liquid–solid fluidized bed results for the studied (ρG =ρL ¼ 0.002), likely due to increased bubble break-up in the
particles before and after CMC addition gave ReL  S;T E1400 ebullated bed, while they remained relatively constant at elevated
(Pjontek and Macchi, 2014) and ReL  S;T E280, respectively. The pressure (ρG =ρL ¼ 0.074). Significant bubble coalescence inhibition
estimated Reynolds number at the terminal free settling velocity (Fig. 6d) again resulted in lower ebullated bed liquid phase
indicated that increasing the liquid viscosity transitioned the parti- holdups when compared to the coalescing system (Fig. 7d).
cles from the Newton (ReL  S;T 4500, inertial forces dominate) to the When comparing Figs. 4c and 6c, it is apparent that the solid
intermediate (0.2oReL  S;T o500) settling flow regimes, where an holdup trends are dependent on both U G =U L and ReL  S . For a
increased liquid viscosity will likely have a greater impact. constant ReL  S and hence constant liquid superficial velocity (shown
740 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

0.55 0.55
coalescing / mixed behaviour
0.5 0.5

Freeboard gas holdup, εG-FB


Bed region gas holdup, εG
0.45 0.45

0.4 0.4

0.35 0.35

0.3 0.3

0.25 0.25

0.2 0.2

0.15 0.15

0.1 0.1
0.6 0.7 0.8 0.9 0.6 0.7 0.8 0.9
Superficial velocity ratio, U G/UL Superficial velocity ratio, UG / UL

0.4 0.6
320

0.36 0.55
ReL-S Bed region liquid holdup, εL
0.32 0.5
Solid holdup, εS

450
0.28 0.45
74

0.24 0.4
ReL-S ρG / ρL
0.002 0.074 ArL-S ReL-S
0.2 0.35 21.0 x 105 320 to 450
104
1.2 x 105 74 to 104

0.16 0.3
0.6 0.7 0.8 0.9 0.6 0.7 0.8 0.9
Superficial velocity ratio, UG/UL Superficial velocity ratio, UG / UL

Fig. 7. Ebullated bed phase holdups for the coalescing (water) and mixed behaviour (0.8 wt% aqueous CMC) systems at varying liquid flow rates and liquid viscosity.

in Fig. 4c), the ebullated bed expanded when increasing U G =U L due presented in this study were on average 23% and 28% greater than
to the increased volumetric gas fraction. However, the opposite trend their associated ebullated bed gas holdups, respectively.
was observed when reducing the liquid superficial velocity for a If bubble characteristics remain constant between the ebullated
constant gas flow rate (shown in Fig. 6c), again increasing U G =U L , as bed and freeboard regions, the bed region gas holdups on a solids-
the lower ReL  S resulted in bed contraction. free basis should be comparable to the freeboard measurements.
The decreasing freeboard gas holdup trends observed with the Fig. 8 compares the solids-free gas holdups in the ebullated bed (i.
0.5 wt% aqueous ethanol prior to CMC addition were expected e., εG =ð1  εS Þ) to the freeboard gas holdups in water and 0.5 wt%
when increasing ReL  S (Fig. 6b) as the rise velocities of the aqueous ethanol using the data from this study and from Pjontek
considerably smaller bubbles are more dependent on the super- and Macchi (2014). The AAREs were 61% and 29% for the coales-
ficial liquid velocity. An improved measurement technique for the cing and coalescence inhibition systems, respectively. For the
gas entrainment in the liquid recirculation would be required to coalescing system, the smaller particles (dSV E1.5 mm) led to
confirm the freeboard gas holdup trend observed following CMC coalesced bubble flow in the ebullated bed due to their particle
addition (Fig. 6b), where a similar reduction at higher liquid flow size and density, where bed contraction at the introduction of gas
rates was initially expected. For the coalescing and mixed beha- was observed. Upon exiting the ebullated bed, bubbles have been
viour systems at atmospheric pressure, the observed bed and visually observed to break-up, likely due to the change in apparent
freeboard gas holdups increase at higher superficial liquid velo- viscosity between the bed and freeboard regions. Fig. 8a demon-
cities (Fig. 7a and b) were likely due to greater shearing on the strates that the solids-free gas holdups at these conditions under-
bubbles when flowing through the distributor plate and ebullated estimated the freeboard gas holdup, in agreement with the bubble
bed, consequently enhancing bubble break-up. break-up when exiting the bed region. The size and density of the
larger particles (dSV E 4 mm) resulted in the dispersed bubble flow
5.3. Relation between bed and freeboard gas holdups regime in water. At these conditions, the solids-free gas holdup
overestimated the freeboard gas holdup, suggesting that bubble
Gas holdup measurements in the LC-FinerSM are limited to the break-up in the ebullated bed due to particle inertia is non-
freeboard region since measurements in the bed region are affected negligible. Similar results were observed in coalescence inhibition
by changes in the catalyst inventory and density, which are not well systems (Fig. 8b), though the solids-free estimate approached the
known while the unit is operational (McKnight et al., 2003). For the freeboard gas holdups when the fluidized bed was sufficiently
coalescing and coalescence inhibition systems, freeboard gas holdups dilute at higher gas holdups.
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 741

Fig. 8. Comparison of solids-free and freeboard gas holdups for (a) water and (b) 0.5 wt% aqueous ethanol. Additional data taken from Pjontek and Macchi (2014).

Fig. 9. Correlated versus experimental gas holdups in the (a) bed and (b) freeboard regions. Additional data taken from Pjontek and Macchi (2014).

5.4. Phase holdup correlations based on the bubble coalescence Table 2


inhibition system Particle settling parameters determined experimentally and using correlations.

Parameter Spheres Cylinders Cylinders


A previous study (Pjontek and Macchi, 2014) compared similar (μL ¼0.001 Pa s) (μL ¼0.004 Pa s)
experimental data obtained using the same experimental system
to ebullated bed correlations available in the literature. It was Experimental n 2.39 2.45 2.64
shown that the correlations struggled to predict the effects of ULT (m/s) 0.33 0.34 0.28
Correlations n 2.44 2.43 2.55
pressure, surfactant addition, and bed expansion or contraction at k 0.83 0.94 0.94
the introduction of gas. The previous discrepancies were primarily ULT1 (m/s) 0.40 0.31 0.29
due to a lack of experimental data at the relevant operating
conditions. Due to the importance of the high gas holdup condi-
tions for hydroprocessing units, the overall phase holdups in
the bed and freeboard regions were thus correlated using the the presence of surfactants, this led to negligible pressure
experimental data for the coalescence inhibition system. App- effects on phase holdups in the ebullated bed. The gas–liquid
licable data from Pjontek and Macchi (2014) was also included density ratio was thus not included in the correlations.
with this study’s results, consequently using spherical and cylind-  Due to the limited range of particle–liquid density ratios studied at
rical particles with Sauter mean diameters of approximately 4 mm
high gas holdups, this dimensionless group was not included.
in 0.5 wt% aqueous ethanol. The correlations were also based on
 The observed high gas holdups resulted from the dispersed
the following considerations:
bubble flow at relatively high superficial gas velocities. Similar
 Gas injection into the liquid using a porous pipe prior to both to the approach of Wilkinson et al. (1992), the provided
correlations are specific to the dispersed bubble flow regime.
phases passing through the distributor plate and subsequent
flow through the ebullated bed contributed to the high gas Table 1 provides the ranges for the dimensionless groups used
holdups by enhancing bubble break-up. When combined with to develop the correlations. Gas holdups in the ebullated bed were
742 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

Fig. 10. Correlated versus experimental solid holdups based on particle settling parameters determined (a) experimentally and (b) from literature correlations. Additional
data taken from Pjontek and Macchi (2014).

correlated using a power-law based on the relevant dimensionless spheres and cylinders was calculated using (Khan and Richardson,
groups. Freeboard gas holdups in a system with significant bubble 1989):
coalescence inhibition and enhanced bubble break-up should not "  0:27 #
4:8  n 0:57 dV
be affected by the particle properties in the ebullated bed. ¼ 0:043Ar L  S 1  1:24 ð9Þ
Experiments at such conditions have shown similar gas holdups n  2:4 dC
between the freeboard region and a bubble column at equal gas Wall effects were estimated based on the relation provided by
and liquid superficial velocities (Pjontek et al., 2014). The free- Khan and Richardson (1989) for spheres and the correlation
board gas holdup correlation was therefore based on the gas– proposed by Chhabra (1995) for cylinders when LP =dp o10:
liquid superficial velocity ratio. It should again be noted that the  0:6
gas and liquid superficial velocities were adjusted to account for dV
spheres : ¼ 1  1:15 ð10Þ
gas entrainment in the liquid recirculation (refer to the method dC
described in Section 5.1.2), thus estimating the actual flow rates in
the ebullated bed. The resulting bed (εG ) and freeboard (εG  FB ) cylinders : ¼ 1  1:33dV =dC ð11Þ
region gas holdups correlations are: The free settling velocity of spherical particles was estimated
 0:94 using the correlation of Turton and Clark (1987), shown to provide
εG UG
¼ 0:62 ReL  S 0:26 ArL  S  0:12 ð7aÞ adequate predictions (Brown and Lawler, 2003). The cylindrical
1  εG UL
terminal free settling velocity was estimated using the Haider and
 1:12 Levenspiel (1989) empirical correlation for isometric non-spherical
εG  FB UG
¼ 0:79 ð7bÞ particles.
1  εG  FB UL
" 0:824  0:412 #  1:214
Fig. 9 compares the predicted and experimental gas
μL 18 0:321
spheres : U LT1 ¼ Ar L 1=3 þ
dV ρL Ar L  S 2=3 Ar L  S 1=3
holdups in the bed and freeboard regions (AARE of 7.0% and
4.7%, AAD of 0.016 and 0.015, respectively). Eqs. (7a) and (7b) ð12Þ
thus provided a satisfactory representation of the gas holdups in  1
the bed and freeboard regions at the selected high gas holdup μL 18 2:335  1:744ϕ
cylinders : U LT1 ¼ Ar L  S 1=3 þ
conditions. dV ρL Ar L  S 2=3 Ar L  S 1=6
Solid holdups were correlated by modifying the well-known ð13Þ
Richardson and Zaki (1954) relationship to include the suggested
Fig. 10 compares the predicted and experimental solids holdups
dimensionless groups for ebullated bed scaling. The Archimedes
using both the experimental and correlated particle settling
number, which is related to the Reynolds number at the terminal
parameters (AARE of 2.0% and 9.7%, AAD of 0.006 and 0.031,
settling velocity of a particle, and the particle–liquid Reynolds are
respectively). The experimentally determined settling parameters
accounted for via the inclusion of the Richardson and Zaki
(Fig. 10a) provided an adequate fit for the solid holdups while the
expression. Since ρS =ρL and ρG =ρL are being excluded based on
correlated parameters (Fig. 10b) resulted in a deviation for the
the prior discussion, the gas–liquid superficial velocity ratio was
aluminum cylinders at a liquid viscosity of 0.001 Pa s. Table 2
included as follows:
demonstrates that the predicted free settling velocity for the
 1=n  0:92 !
UL UG previous conditions underestimated the experimentally deter-
1  εS ¼ 1 þ 0:22 ð8Þ mined value, in agreement with the observed solid holdup devia-
kU LT1 UL
tion (shown in Fig. 10b). It is consequently recommended to use
Coefficients for the U G =U L dimensionless group were deter- experimentally determined liquid–solid fluidized bed settling
mined using the experimentally determined terminal settling parameters when possible. The proposed correlations thus
velocities (U LT ¼ k U LT1 ) and n index from the liquid–solid flui- appeared to adequately represent the bed and freeboard gas
dized beds. Nonetheless, empirical correlations for U LT and n were holdups as well as solid holdups for a system operating at high
compared to the experimentally determined values (refer to gas holdups and that satisfies the discussed reactor geometric and
Table 2) to investigate the robustness of Eq. (8). The n index for bubble coalescence characteristics.
D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744 743

6. Conclusions n index for bed expansion correlation


P pressure (Pa)
Dynamic similitude was assumed when matching five dimen-  ΔP dynamic pressure drop (Pa)
sionless groups, at equal bubble coalescence behaviour based on a ReL  S particle–liquid Reynolds number, ReL  S ¼ U L ρL dSV =μL
binary consideration (e.g., coalescing or coalescence inhibition), ReL  S;T particle–liquid Reynolds number based on terminal free
and relaxing important geometric considerations (e.g., gas injec- settling velocity and accounting for wall effects,
tion method and gas–liquid distribution) for systems at high gas ReL  S;T ¼ U LT ρL dSV =μL
holdups. This approach was tested by comparing the overall phase T temperature (1C)
holdups for separate cylindrical particles (dSV of 1.6 and 3.9 mm) at UG, UL gas and liquid superficial velocities (m/s)
matching dimensionless groups, where results were comparable U LT terminal settling velocity of a particle, accounting for
when bubble coalescence was consistently and sufficiently inhib- wall effects (m/s)
ited. Unfortunately, the comparison could not be carried out in a U LT1 terminal free settling velocity of a particle (m/s)
coalescing system as CMC addition to water resulted in some vB bubble rise velocity (m/s)
surface-active characteristics that affected the overall phase hold- Δz vertical distance between differential pressure taps (m)
ups, particularly at low gas velocities.
The combined effects of enhanced bubble break-up (i.e., gas Greek symbols
injection, gas–liquid distribution method, and/or elevated pres-
sure) and significant bubble coalescence inhibition (i.e., surfactant γG  L gas–liquid surface tension (N/m)
addition) were required to achieve the desired high gas holdup εG , εL , εS gas, liquid and solid holdups in the bed region
conditions. The effects of increased liquid viscosity, varying super- εG  FB freeboard gas holdup
ficial gas velocity and varying superficial liquid velocity were μG , μL gas and liquid dynamic viscosity (Pa s)
therefore studied at relevant fluid dynamic conditions for indus- ρG , ρL , ρS gas, liquid and solid densities (kg/m3)
trial hydroprocessors. When increasing the liquid viscosity in the φ sphericity
0.5 wt% aqueous ethanol, a fraction of the gas was entrained in the
liquid recirculation due to inadequate foam dissipation at the free-
Subscripts
surface. The increased liquid viscosity mainly reduced the solids
holdups, after accounting for the gas recirculation and adjusting
FB freeboard
the gas and liquid flow rates. Gas entrainments of up to approxi-
G gas
mately 20 vol% at the highest studied gas flow rates resulted in
L liquid
similar gas holdups when compared to previous industrial
P particle
measurements.
S solid
For the coalescing and coalescence inhibition systems, free-
board gas holdups were on average 23% and 28% greater than bed
region gas holdups, respectively. When attempting to estimate
freeboard gas holdups based on a solids-free basis in the ebullated
bed for water and 0.5 wt% aqueous ethanol, the AAREs were 61% Acknowledgments
and 29%, respectively, where deviations were mainly due to
enhanced bubble break-up or coalescence due to bubble–particle The authors are grateful to Dr. Jules Thibault for allowing the
interactions. The proposed bed and freeboard gas holdup correla- use of the Anton Paar Physica MCR 301 Rheometer. The authors
tions provided an adequate fit at the selected high gas holdup would like to acknowledge the Natural Sciences and Engineering
conditions (AARE of 8.9% and 5.4%, respectively). Solids holdups Research Council of Canada, the Canadian Foundation for Innova-
were correlated based on a modified Richardson and Zaki (1954) tion, the Ontario Innovation Trust, and Syncrude Canada Ltd. for
expression which also provided an acceptable fit (AARE of 2.0%), financial support.
particularly when the particle terminal settling velocity and n
index were determined from experimental liquid–solid fluidized
bed results. References

Behkish, A., Lemoine, R., Oukaci, R., Morsi, B.I., 2006. Novel correlations for gas
holdup in large-scale slurry bubble column reactors operating under elevated
Nomenclature pressures and temperatures. Chem. Eng. J. 115, 157–171. http://dx.doi.org/
10.1016/j.cej.2005.10.006.
AARE average absolute relative error Brown, P.P., Lawler, D.F., 2003. Sphere drag and settling velocity revisited. J. Environ.
3 Eng. 129, 222–231. http://dx.doi.org/10.1061/(ASCE)0733-9372(2003)129:3
Ar L  S particle–liquid Archimedes number, Ar L  S ¼ ρL dV ρS  (222).
ρL Þ g=μL 2
Chhabra, R.P., 1995. Wall effects on free-settling velocity of non-spherical particles
CMC carboxymethyl cellulose in viscous media in cylindrical tubes. Powder Technol. 85, 83–90. http://dx.doi.
org/10.1016/0032-5910(95)03012-X.
dC column inner diameter (m) Dargar, P., Macchi, A., 2006. Effect of surface-active agents on the phase holdups of
dP particle diameter (m) three-phase fluidized beds. Chem. Eng. Process. Process Intensif. 45, 764–772.
dSV Sauter mean diameter (m) http://dx.doi.org/10.1016/j.cep.2006.03.004.
Davies, R.M., Taylor, G., 1950. The mechanics of large bubbles rising through
dV volume equivalent diameter
 (m) 2 extended liquids and through liquids in tubes. Proc. R. Soc. A Math. Phys. Eng.
Eo Eötvös number, Eo ¼ g ρL  ρG dP =γ G  L Sci. 200, 375–390. http://dx.doi.org/10.1098/rspa.1950.0023.
2
g gravitational acceleration (m/s ) Fan, L.-S., Bavarian, F., Gorowara, R., Kreischer, B.E., Buttke, R.D., Peck, L.B., 1987.
Hydrodynamics of gas–liquid–solid fluidization under high gas hold-up condi-
hB fluidized bed height (m)
tions. Powder Technol. 53, 285–293. http://dx.doi.org/10.1016/0032-5910(87)
hC column height (m) 80101-8.
k wall effect for bed expansion correlation Fan, L.-S., Tsuchiya, K., 1990. Bubble Wake Dynamics in Liquids and Liquid–Solid
LP particle length (m) Suspensions. Butterworth-Heinemann, Stoneham.
Fan, L.-S., Yang, G.Q., Lee, D.J., Tsuchiya, K., Luo, X., 1999. Some aspects of high-
m mass of the particles
 (kg) pressure phenomena of bubbles in liquids and liquid–solid suspensions. Chem.
M M-group, M ¼ g ρL  ρG μL 4 =ρL 2 γ G  L 3 Eng. Sci. 54, 4681–4709. http://dx.doi.org/10.1016/S0009-2509(99)00348-6.
744 D. Pjontek et al. / Chemical Engineering Science 126 (2015) 730–744

Gorowara, R.L., Fan, L.-S., Gorowarat, R.L., 1990. Effect of surfactants on three-phase Rana, M.S., Sámano, V., Ancheyta, J., Diaz, J.A.I., 2007. A review of recent advances
fluidized bed hydrodynamics. Ind. Eng. Chem. Res. 29, 882–891. http://dx.doi. on process technologies for upgrading of heavy oils and residua. Fuel 86,
org/10.1021/ie00101a025. 1216–1231. http://dx.doi.org/10.1016/j.fuel.2006.08.004.
Haider, A., Levenspiel, O., 1989. Drag coefficient and terminal velocity of spherical Richardson, J.F., Zaki, W.N., 1954. Sedimentation and fluidisation, Part 1. Trans. Inst.
and nonspherical particles. Powder Technol. 58, 63–70. http://dx.doi.org/ Chem. Eng. 32, 357–362.
10.1016/0032-5910(89)80008-7. Ruiz, R.S., Alonso, F., Ancheyta, J., 2004. Effect of high pressure operation on overall
Hughmark, G.A., 1967. Holdup and mass transfer in bubble columns. Ind. Eng. phase holdups in ebullated-bed reactors. Catal. Today 98, 265–271. http://dx.
Chem. Process Des. Dev. 6, 218–220. http://dx.doi.org/10.1021/i260022a011. doi.org/10.1016/j.cattod.2004.07.039.
Ishibashi, H., Onozaki, M., Kobayashi, M., Hayashi, J.-i., Itoh, H., Chiba, T., 2001. Gas Ruiz, R.S., Alonso, F., Ancheyta, J., 2005. Pressure and temperature effects on the
holdup in slurry bubble column reactors of a 150t/d coal liquefaction pilot plant hydrodynamic characteristics of ebullated-bed systems. Catal. Today 109,
process. Fuel 80, 655–664. http://dx.doi.org/10.1016/S0016-2361(00)00131-9. 205–213. http://dx.doi.org/10.1016/j.cattod.2005.08.019.
Kantarci, N., Borak, F., Ulgen, K.O., 2005. Bubble column reactors. Process Biochem. Safoniuk, M., Grace, J.R., Hackman, L., McKnight, C.A., 1999. Use of dimensional
40, 2263–2283. http://dx.doi.org/10.1016/j.procbio.2004.10.004. similitude for scale-up of hydrodynamics in three-phase fluidized beds. Chem.
Kelkar, B.G., Godbole, S.P., Honath, M.F., Shah, Y.T., Carr, N.L., Deckwer, W., 1983. Eng. Sci. 54, 4961–4966. http://dx.doi.org/10.1016/S0009-2509(99)00218-3.
Effect of addition of alcohols on gas holdup and backmixing in bubble columns. Sagert, N.H., Quinn, M.J., 1977. Influence of high-pressure gases on the stability of
AIChE J. 29, 361–369. http://dx.doi.org/10.1002/aic.690290303. thin aqueous films. J. Colloid Interface Sci. 61, 279–286. http://dx.doi.org/
Khan, A.R., Richardson, J.F., 1989. Fluid–particle interactions and flow character- 10.1016/0021-9797(77)90391-5.
istics of fluidized beds and settling suspensions of spherical particles. Chem. Sagert, N.H., Quinn, M.J., 1978. Surface viscosities at high pressure gas–liquid
Eng. Commun. 78, 111–130. http://dx.doi.org/10.1080/00986448908940189. interfaces. J. Colloid Interface Sci. 65, 415–422. http://dx.doi.org/10.1016/0021-
Krishna, R., Wilkinson, P.M., Van Dierendonck, L.L., 1991. A model for gas holdup in
9797(78)90092-9.
bubble columns incorporating the influence of gas density on flow regime
Sanchez, J., Ruiz, R.S., Alonso, F., Ancheyta, J., Sánchez, J.L., 2008. Evaluation of the
transitions. Chem. Eng. Sci. 46, 2491–2496. http://dx.doi.org/10.1016/0009-
hydrodynamics of high-pressure ebullated beds based on dimensional simili-
2509(91)80042-W.
tude. Catal. Today 130, 519–526. http://dx.doi.org/10.1016/j.cattod.2007.10.002.
Larachi, F., Belfares, L., Iliuta, I., Grandjean, B.P.a., 2001. Three-phase fluidization
Shah, Y.T., Joseph, S., Smith, D.N., Ruether, J.A., 1985. On the behavior of the gas
macroscopic hydrodynamics revisited. Ind. Eng. Chem. Res. 40, 993–1008. http:
phase in a bubble column with ethanol–water mixtures. Ind. Eng. Chem.
//dx.doi.org/10.1021/ie0006864.
Process Des. Dev. 24, 1140–1148. http://dx.doi.org/10.1021/i200031a041.
Luo, X., Jiang, P., Fan, L.-S., 1997. High-pressure three-phase fluidization: hydro-
Shah, Y.T., Kelkar, B.G., Godbole, S.P., Deckwer, W.-D.D., 1982. Design parameters
dynamics and heat transfer. AIChE J. 43, 2432–2445. http://dx.doi.org/10.1002/
estimations for bubble column reactors. AIChE J. 28, 353–379. http://dx.doi.org/
aic.690431007.
10.1002/aic.690280302.
Luo, X., Lee, D.J., Lau, R., Yang, G., Fan, L.-S., 1999. Maximum stable bubble size and
Shaikh, A., Al-Dahhan, M., 2010. A new methodology for hydrodynamic similarity in
gas holdup in high-pressure slurry bubble columns. AIChE J. 45, 665–680. http:
//dx.doi.org/10.1002/aic.690450402. bubble columns. Can. J. Chem. Eng. 88, 503–517. http://dx.doi.org/10.1002/
Macchi, A., Bi, H., Grace, J.R., McKnight, C.A., Hackman, L., 2001. Dimensional cjce.20357.
hydrodynamic similitude in three-phase fluidized beds. Chem. Eng. Sci. 56, Shaikh, A., Al-Dahhan, M., 2013. Scale-up of bubble column reactors: a review of
6039–6045. http://dx.doi.org/10.1016/S0009-2509(01)00207-X. current state-of-the-art. Ind. Eng. Chem. Res. 52, 8091–8108. http://dx.doi.org/
Macchi, A., Bi, H., Grace, J.R., McKnight, C.A., Hackman, L., 2003. Effect of gas density 10.1021/ie302080m.
on the hydrodynamics of bubble columns and three-phase fluidized beds. Can. Song, G.-H., Bavarian, F., Fan, H.-S., Buttke, R.D., Peck, L.B., 1989. Hydrodynamics of
J. Chem. Eng. 81, 846–852. http://dx.doi.org/10.1002/cjce.5450810368. three-phase fluidized bed containing cylindrical hydrotreating catalysts. Can. J.
McKnight, C.A., Hackman, L.P., Grace, J.R., Macchi, A., Kiel, D., Tyler, J., 2003. Fluid Chem. Eng., 67; , pp. 265–275. http://dx.doi.org/10.1002/cjce.5450670213.
dynamic studies in support of an industrial three-phase fluidized bed hydro- Tarmy, B.L., Chang, M., Coulaloglou, C.A., Ponzi, P.R., 1984. Three phase hydro-
processor. Can. J. Chem. Eng. 81, 338–350. http://dx.doi.org/10.1002/ dynamic characteristics of the eds coal liquefaction reactors: their development
cjce.5450810302. and use in reactor scaleup. Inst. Chem. Eng. Symp. Ser. 87, 303–317.
Pjontek, D., Macchi, A., 2014. Hydrodynamic comparison of spherical and cylind- Tarmy, B.L., Coulaloglou, C.A., 1992. Alpha-omega and beyond industrial view of
rical particles in a gas–liquid–solid fluidized bed at elevated pressure and high gas/liquid/solid reactor development. Chem. Eng. Sci. 47, 3231–3246. http://dx.
gas holdup conditions. Powder Technol. 253, 657–676. http://dx.doi.org/ doi.org/10.1016/0009-2509(92)85031-6.
10.1016/j.powtec.2013.12.030. Turton, R., Clark, N., 1987. An explicit relationship to predict spherical particle
Pjontek, D., Parisien, V., Macchi, A., 2014. Bubble characteristics measured using a terminal velocity. Powder Technol. 53, 127–129.
monofibre optical probe in a bubble column and freeboard region under high Wilkinson, P.M., Spek, A.P., van Dierendonck, L.L., 1992. Design parameters estima-
gas holdup conditions. Chem. Eng. Sci. 111, 153–169. http://dx.doi.org/10.1016/j. tion for scale-up of high-pressure bubble columns. AIChE J. 38, 544–554. http:
ces.2014.02.024. //dx.doi.org/10.1002/aic.690380408.

Das könnte Ihnen auch gefallen