You are on page 1of 9

Chemical Engineering Science 123 (2015) 620–628

Contents lists available at ScienceDirect

Chemical Engineering Science


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

How methylhydroxyethylcellulose (MHEC) influences drying


in porous media
A.P.A. Faiyas a,b, S.J.F. Erich b,c,n, M. van Soestbergen a,b, H.P. Huinink b,
O.C.G. Adan b,c, T.G. Nijland c
a
Materials innovation institute, Mekelweg 2, P.O.Box 5008, 2628 CD Delft, The Netherlands
b
Department of Applied Physics, Eindhoven University of Technology, Den Dolech 2, P.O.Box 513, 5600 MB, Eindhoven, The Netherlands
c
Netherlands Organization for Applied Scientific Research (TNO), P.O.Box 49, 2600 AA, Delft, The Netherlands

H I G H L I G H T S

 Transition from homogeneous to front receding drying due to presence of MHEC.


 Presence of MHEC unexpectedly reduces evaporation.
 The observed change in drying behavior in presence of MHEC is mainly a result of viscosity.
 In presence of MHEC a clear minimum moisture diffusivity is observed.

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

Article history: This article presents both an experimental as well as a theoretical study on the effect of MethylHy-
Received 31 August 2014 droxyEthylCellulose (MHEC) on drying in porous materials using Nuclear Magnetic Resonance Imaging
Received in revised form (NMR). MHEC, a water soluble polymer, is normally added to glue mortars as a water retention agent in
24 November 2014
order to improve the drying by increasing the open time. However, the exact processes that determine
Accepted 25 November 2014
Available online 3 December 2014
the drying rate of a glue mortar are unknown. In this study, we therefore focus on investigating the
drying of a Fired Clay Brick (FCB) saturated with an aqueous MHEC solution. By using a FCB as a model
Keywords: system, the influence of hydration and changing of pore sizes due to hydration can be removed. The
Viscous fluid performed NMR experiments show a transition from homogeneous drying for a water saturated FCB
Porous media
toward an inhomogeneous (front receding) drying behavior for FCB saturated with increasing MHEC
Transport processes
concentration. The capillary number (Ca) indicates a homogeneous drying for water saturated brick
Diffusion
Drying (Ca«1) and an inhomogeneous drying (Ca»1) in case of saturated brick with more than 1.5 wt% MHEC.
NMR Analysis of capillary number shows that the main parameter determining the capillary numbers are
viscosity, surface tension, contact angle and evaporation rate. Among this viscosity change match with
change in capillary number and therefore viscosity has a major influence on the drying behavior and
rate. Based on the measured profiles, the moisture diffusivity is calculated. Using an empirical equation,
the moisture diffusivity and NMR profiles were fitted to obtain the key parameters. From these analyses
we conclude that 1) the moisture diffusivity scales with viscosity, which is the main parameter in shift in
drying behavior. 2) Surprisingly, the presence of MHEC reduced the evaporation, which is unexpected
since evaporation of MHEC solutions, for instance present in capillaries do not exhibit this behavior. 3) A
minimum moisture diffusivity is found which shows to correlate to the evaporation process, which
indicates that this might be the result of a vapor dominated diffusion flux.
& 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Methylhydroxyethylcellulose (MHEC) is widely used as water


n
retention agent in cementitious materials. The major applications
Correspondence to: Transport in Permeable Media, Department of Applied
are tile adhesives, wall renders, self-leveling underlayments, floor
Physics, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The
Netherlands. Tel.: þ31 40 2473830; fax: þ31 40 2432598. screeds and water-proofing membranes. However, the way MHEC
E-mail address: s.j.f.erich@tue.nl (S.J.F. Erich). retains water, and as such influences the open time (OT) of cement

http://dx.doi.org/10.1016/j.ces.2014.11.054
0009-2509/& 2014 Elsevier Ltd. All rights reserved.
A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628 621

mortar, is not fully understood. Open time is the time during which 2. Experimental
tiles can be applied with sufficiently good adhesion. This open time
can be modified by adding a small amount (1–2 wt%) of water 2.1. Methylhydroxyethylcellulose (MHEC)
soluble polymer, which might be caused by a change in drying
behavior of cementitious glue mortar. Methylhydroxyethylcellulose (MHEC) is a water soluble cellulose
Drying of porous media is a complex process (Prat, 1993; Metzger ether, synthesized from cellulose by substituting methyl and ethy-
and Tsotsas, 2005; Yiotis et al., 2005; Kopinga and Pel, 1994; Prat, lene oxide groups by an etherification process. A commercial grade
2002; Le Bray and Prat, 1999). In most studies, drying in porous MHEC (Tylose MHS 6000 P6) provided by SE Tylose GmbH & Co.KG,
media focuses on low-viscous fluids (e.g. water). In most cases, a Germany, with a degree of substitution (DSME, ME¼Methyl) of
porous material dries homogeneously with a constant drying rate 1.3 and a molar degree of substitution (MSHE, HE¼Hydroxyethyl)
until a critical moisture content is reached (Prat et al., 2012). In the of 0.3 was used in our experiments. The MHEC solutions were
first stage, drying is dominated by the external mass transfer (e.g. prepared by heating 200 ml of water to 80 1C allowing the dissolu-
evaporation). At the critical moisture content, the fluid path is no tion of MHEC. The solution was homogenized by stirring the solution
longer continuous and drying no longer homogeneous. From this with a magnetic stirrer at 500 rpm for two days. Finally, the solution
point onward the drying process takes place by transport of vapor was cooled to room temperature.
and/or liquid films. In this stage, drying is dominated by internal The dynamic viscosity of MHEC solutions were measured by
mass transfer (e.g. internal vapor transport). Different transport TA-Instruments AR-1000 rheometer. Fig. 1 shows the dynamic
processes inside porous materials may result in similar drying viscosity as function of MHEC concentration. Interesting feature
behavior. In the latter case, the drying behavior is almost similar to from Fig. 1 is that the viscosity value of 4.3 wt% MHEC varies four
that observed in case of drying of squared or other shaped capillaries orders of magnitude compared to 0 wt% MHEC.
in which liquid films are present in the corners of the capillaries The surface tension as a function of MHEC concentration is
(Chauvet et al., 2009; Prat, 2007). plotted in Fig. 1. Conventional pendant-drop technique was used
Resolving the exact drying processes in a porous media is to measure the surface tension of the MHEC solutions (DSA 100
difficult, because in-situ measurement of drying materials is not Drop Shape Analyzer). The surface tension of MHEC solutions
straightforward. As a consequence, many studies focus on simulat- was determined by fitting the drop shape to the Young Laplace
ing drying behavior, for example, through pore network models equation which relates the interfacial tension to the drop shape.
(PNM). Metzger et al. (2008) (Metzger and Tsotsas, 2008) simulated Fig. 1 shows a decrease in surface tension with increase in MHEC
a 3D pore network both with and without viscous forces (water) concentration. The most pronounced decrease is observed at
and studied the viscous stabilization of the drying front and its low MHEC concentrations.
effect on the drying rate. These models show that the viscous forces The contact angle for different MHEC concentration is deter-
of water play a role in the intermediate stages of drying, by shifting mined both on silica glass and on MHEC film and is also given in
the profile from a homogeneous to inhomogeneous drying behavior Fig. 1. The contact angle was determined using a data physics OCA-
(the second stage). As far as we know, no experiments have been 20 contact angle instrument. The contact angle of MHEC solution
performed with high viscous fluids in porous materials. Only few by the sessile drop method. In both cases, an increase in contact
investigations focus on the drying of porous media and investigate angle is observed with increase in MHEC concentration.
the effect of changes in contact angle, for example, hydrophobic or
hydrophilic surfaces (Prat, 2011). In the latter case, it was shown
that hydrophobic surfaces having a contact angle above a critical
contact (e.g. 1201) exhibit a flat traveling drying front and that
drying is independent of the contact angle.
As indicated, measuring transport processes inside a real 3D
porous material is challenging, because porous media are in
general non transparent. Fortunately, some 2D model porous
systems consisting of glass beads exist, allowing imaging the
drying process using a camera (Lu et al., 1994). To measure fluid
distributions inside 3D porous materials techniques such as
synchrotron X-ray tomography (Shokri and Sahimi, 2012) and
Nuclear Magnetic Resonance (NMR) imaging seem to be the only
available techniques probing the moisture in a non-invasive and
non-destructive manner with a sufficiently high spatial and time
resolution (Pel et al., 1996).
This study focusses on the experimental investigation of drying
of porous materials saturated with methylhydroxyethylcellulose
(MHEC) solution and aims to understand the effect of viscosity,
surface tension, and contact angle on the drying process by
measuring the drying process with NMR. Since glue mortars are
hydrating, the total water content of the mortar is reduced by both
hydration reaction and evaporation. To uncouple hydration from
evaporation induced transport processes, Fired Clay Brick (FCB)
was chosen as a non-reactive medium.
The article will start with the experimental details of the
investigation followed by a theory section providing a basis for
interpreting the experimental results (presented in the subsequent
section). To understand the drying processes, a model is used that Fig. 1. In this figure a) viscosity, b) surface tension and c) contact angle of MHEC
describes the transport processes, which is validated against the solution are plotted as function of MHEC concentration.
measured profiles.
622 A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628

2.2. Fired clay brick (FCB) as a model porous media

Fired clay brick (FCB) is used as a model porous media to


understand the influence of MHEC on the drying. The brick has an
average porosity of 0.31 m3m  3 with a permeability value of
9.6  10–14 m2 (Hall and Hoff, 2012). The average pore size is
approximately 5 mm as determined by Mercury Intrusion Porosi-
metry (MIP). The cumulative pore volume against pore diameter
for FCB is presented in Fig. 2; the pore size distribution depends on
the pore geometry, as MIP actually probes the diameter of the pore
entrance.

2.3. NMR imaging

Nuclei with a net magnetic moment will start to precess when


placed in a magnetic field with a resonance frequency, called the
Larmor frequency given by,
γ
f ¼ B0 ; ð1Þ
2π Fig. 3. Schematic diagram of NMR set-up.
where γ =2π is the gyromagnetic ratio of the nuclei (42.58 MHz/T
for 1H), and B the applied magnetic field strength. The constant
of 2.5 s and with eight number of averages at different sample
magnetic field can be replaced by a spatially varying magnetic
positions divided by the reference sample (containing water) of
field B,
equal volume. The profiles are normalized with the porosity of the
B ¼ B0 þ ðG U xÞ; ð2Þ medium.
where G represents the magnetic field gradient and x the position.
This enables in-situ monitoring of moisture content at different 2.5. Drying experiments
positions in the sample due to the spatial variations in the
resonance frequency. One dimensional NMR drying experiments were performed on
The sum of all individual magnetic moments of the spins is the FCB samples saturated both with and without MHEC. Cylindrical
called net macroscopic magnetization, and can be manipulated by samples were drilled with a diameter of 18 mm and a length of
applying an oscillating magnetic field at the resonance frequency 10 mm. One sample was vacuum saturated with pure water and
of the nuclei. The intensity of the resulting spin echo signal is the rest were vacuum saturated with different concentrations of
proportional to the density of the magnetic moment and thus the MHEC in aqueous solution, namely 0.4 wt%, 1.3 wt%, 2.1 wt% and
hydrogen density. The hydrogen density is directly connected to 4.3 wt%. Each sample was vacuum saturated for three days to
the local water content. ensure full saturation. The drying was performed by an air flow
of 1 l min  1 introduced vertically in the middle of the sample
2.4. NMR setup producing an almost zero relative humidity on top of the sample.
All other sides of the samples were closed by a cylindrical
The experiments were performed with a home built NMR setup Teflon tube.
with a static magnetic field of 0.7 T, resulting in a resonance
frequency of 31 MHz and a gradient of 400 mT/m. In this config-
uration a spatial resolution of 0.8 mm was obtained. Fig. 3 shows 3. Theory
the schematic diagram of the NMR set up designed to measure
moisture profile of cylindrical sample during drying. The step This section aims to provide the theoretical basis for the
motor is used to position the sample inside the NMR set-up. calculation and interpretation of the moisture profiles based on
Moisture profiles were obtained with a Hahn spin echo sequence the material parameters, boundary conditions and the obtained
with an echo time of 150 ms, window width 100 ms, repetition time moisture diffusivity determined from the moisture profiles.

0.08 3.1. Transport equation and moisture diffusivity


0.20
0.07
One-dimensional moisture transport in porous media under
Incremental pore volume (ml/g)

cumulative pore volume (ml/g)

0.06 isothermal conditions can be described by the following conserva-


0.15 tion equation:
0.05
∂Θ ∂
þ qðΘÞ ¼ 0; ð3Þ
0.04
0.10
∂t ∂x
0.03 in which Θ [m3/m3] represents the moisture content and the
moisture flux q [m3/m2s] consists of vapor transport and pressure
0.02 0.05 driven Darcy flow, given by:
0.01   Dv ∂ρv K ∂pc
q Θ ¼ þ ; ð4Þ
ρw ∂x η ∂x
0.00 0.00
0.1 1 10 100
where Dv [m2/s] represents the vapor diffusion coefficient, ρv [kg/m3]
Mean Diameter (µm)
the vapor density, ρw [kg/m3] the density of water, K [m2] the
Fig. 2. Cumulative pore volume vs. pore diameter for FCB from MIP. permeability, η [Pa s] the viscosity and pc [Pa] the capillary pressure.
A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628 623

Phenomenological, a moisture transport equation can be described constant for each sample. In essence, Eq. (11) describes the ratio
by a non linear diffusion equation: between viscous forces and capillary forces in a specific porous
∂Θ ∂ ∂Θ material that is drying at the surface. When Ca «1, capillary forces
þ DðΘÞ ¼ 0; ð5Þ dominate over the viscous forces. In case Ca»1, the capillary forces
∂t ∂x ∂x
(present due to the evaporation process at the surface) no longer
where Θ is the moisture content and DðΘÞ the moisture diffusivity, surpass the viscous forces in the permeable material. Looking at the
which is depending on the local moisture content. The moisture properties of the MHEC fluid in Fig. 1, we conclude that the viscosity
diffusivity now covers both the vapor transport as well as the Darcy of MHEC changes several orders of magnitude, whereas the surface
flow. Determination of the moisture diffusivity DðΘÞ in porous media, tension and contact angle only influence the capillary number by an
incorporating the above two processes, using moisture concentration order of magnitude.
profiles can be achieved by using the following equation (Landman One can re-write the Eq. (11) by using Darcy's law as,
et al., 2001; Pel et al., 1996)
R x0   L
∂Θ=∂t dx Ca ¼ ; ð12Þ
DðΘðx0 ÞÞ ¼ L   ; ð6Þ ξ
∂Θ=∂x x0
where L is the size of the sample and ξ represents the width of the
where L denotes the length of the sample and the position at the drying front. By using this equation, we can directly relate the
back of the sample. Physically, the moisture diffusivity contains two value of capillary number to the drying behavior. In case Ca«1, the
processes, vapor diffusion and liquid flow, and is given by: width of the front is much larger than the size of the sample,
  Dv ∂ρv K ∂pc which represents homogeneous drying of the material. If Ca» 1, the
D Θ ¼  : ð7Þ
ρw ∂ Θ η ∂ Θ width of the front is much smaller than the size of the sample,
which corresponds to inhomogeneous drying characterized by a
The vapor density ρv can be related to Pc through the Kelvin
receding front (Lenormand. et al., 1988).
equation:
M w pc
ρv ¼ ρ0 e  ρw RT ; ð8Þ
4. Result and discussion
where ρ0 [22 g/cm3] is the vapor density of water, ρw [kg/m3] the
density of water, M w [18 g/mol] the molar mass of water, T [K] the 4.1. Drying profiles
temperature and R [8.314 J/mol K] the gas constant. To calculate
both the vapor pressure and liquid flow, the capillary pressure of With NMR imaging, the moisture distribution in drying fired
the material needs to be calculated, which can be achieved from clay brick samples was obtained. Fig. 4 shows the moisture
mercury intrusion porosimetry and fitted with the Van Genugten distribution of fired clay brick for different cellulose concentra-
equation (Genuchten, 1980): tions at different times during drying. The top and bottom side of
  m !m1 the samples are located on left and right side of the profiles,
θmax m1
pc ¼ p0c 1 ; ð9Þ respectively.
θ To investigate the difference in drying behaviour, the cellulose
where p0c is the maximum capillary pressure (for our FCB 258 kPa), ether concentration within the FCB is varied. Fig. 4(a) shows the
θmax (for our FCB 0.31 m3/m3) is the moisture content when fully drying in case of pure water. The first curve (black) represents the
saturated, and m [1.96] a constant. profile after 20 min of averaging. The profiles are acquired and
To describe the transport also the evaporation process at the plotted every 20 min. After seven profiles, the profile is plotted
boundary should be taken into account. The vapor flux is deter- every 40 min to visualise the drying process. Homogeneous drying
mined by the vapor diffusion (Dair Þ induced by a vapor density is observed, as indicated by the vertical arrow and continues till a
difference (Δρv Þ over a surface layer with a thicknessðδÞ with dry critical saturation, θc, reached at E0.08 m3 m  3, indicated by the
air at the outer part of this layer (ρδv  0Þ and a sample surface horizontal arrow. After the critical saturation has been reached, a
vapor density of ρsv . receding drying front develops and moves into the brick.
The moisture profile for the sample having 0.4 wt% (Fig. 4(b)) of
Δρv ¼ ρsv  ρδv  ρsv ; ð10Þ MHEC shows homogeneous drying behavior similar to the
The flux over this surface area can be calculated by j ¼ drying observed in pure water saturated brick. However, with
ðDair Δρv =δÞ ¼ ðDair ρsv =δÞ. The vapor pressure at the direct surface increasing concentration, as seen in Fig. 4(c) with 1.3 wt% MHEC,
of the sample depends on the saturation level of the material at the the drying behavior becomes more inhomogeneous. A receding
surface, and can be calculated by the Kelvin equation (Eq. 8). drying front is observed from the start of the drying process, as
indicated by the horizontal arrow. This shows that the capillary
3.2. Capillary number forces are not large enough to sustain sufficient flow. However, at
the back of the sample, the moisture content still decreases quite
One can characterize the drying inside a porous material using homogeneously, indicating that the capillary forces are still driving
the capillary number (Ca). The capillary number is the ratio the flow on smaller length scales. The moisture profiles shown in
between the viscous force and the capillary force is given by, Fig. 4(d) (4.3 wt%) also show an inhomogeneous drying behavior. A
moving front is observed as indicated by the horizontal arrow.
Lnνηr
Ca  ; ð11Þ Additionally, one can observe a gradient in moisture content (with a
2 K γ cos θ
typical length scale of 7 mm) at the deeper layers (indicated by the
where n [-] represents the porosity, ν [m/s] the fluid velocity, η [Pa s] second arrow). Additionally, at the bottom, more homogenous
the viscosity, r [m] the pore radius, K [m2] the permeability, γ [N/m] drying is visible (indicated by the vertical arrow). A similar drying
the surface tension and θ [1] the contact angle. This equation shows behavior was observed for the samples with 2.1 wt% MHEC solution.
that the drying behavior is mainly determined by evaporation rate In order to characterize the drying, the total moisture content is
(which determines the fluid velocity ν), viscosity, the permeability of plotted in Fig. 5 against time for different wt.% of MHEC. The total
the material, surface tension and the contact angle of the pore liquid, moisture content is obtained by integrating the moisture profiles.
considering that the pore radius, sample size and porosity are The symbols plotted in this figure represent the measured data,
624 A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628

Fig. 4. Moisture profile of the sample as function of position during drying. a) 0 wt% plotted every 20 min, b) 0.4 wt% c) 1.3 wt% and d) 4.3 wt% profiles plotted every 30 min.

4.2. Understanding drying behavior

This section is focused on determining the key parameters that


determine the shift from homogeneous drying to front drying. This
is achieved by investigating the dependence of the capillary
number on the addition of MHEC. Subsequently, the drying
behavior is investigated, by determining the moisture diffusivity
as a function of moisture content for the different MHEC saturated
bricks. An empirical diffusivity equation is used to fit both the
calculated moisture diffusivity as a function of moisture content
and the measured NMR profiles. This enables determining the
most influential parameters and processes.

4.2.1. Capillary number


The parameters determining the capillary number from Eq. (11)
Fig. 5. Total volume of water present in the brick samples for different concentra- are given in Table 1. The pore velocity was determined from the
tions of MHEC as function of time. The open symbols represent the measurement initial slope of the total moisture content curve divided by the
data, the lines are fits from the drying model. porosity of the media at t¼0 in Fig. 5. The pore radius was
determined from the MIP measurement. The permeability was
determined from measuring the flow through a brick sample with
while the continuous line represents a model fit, which will be a water column on top. The dynamic viscosity, surface tension and
discussed in the next sections of this article. Different drying contact angle as function of MHEC concentration are given in
stages can be distinguished. In the first period for 0 wt% MHEC, the Fig. 1. In the contact angle measurements, both glass and MHEC
drying curve is decreasing linearly. In the last period of drying, the film substrate have been measured, because during drying, a
moisture decrease is non-linear and the evaporation rate drops. MHEC film may possibly form on the internal pore wall, influen-
This corresponds to the period in which a receding front moves cing the contact angle. Note the difference in contact angle
inside the material and is dominated by internal evaporation between glass and MHEC film, and the change of contact angle
process and subsequent diffusion. as a function of the MHEC concentration. In case of MHEC film, the
A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628 625

Table 1
Experimental parameters used to calculate the capillary number.

Sample Parameters

Sample length, Porosity, Pore velocity, Dynamic viscosity, μ Pore radius, Permeability, Surface tension, Contact angle, θ Capillary number,
(mm) n (m3m  3) v (μm/s) (Pas) r (μm) K (m2) γ (mN/m) (1) Ca (-)

0 wt% 10 0.31 1.58 1  10  3 5 9.6  10–14 70 16 1.9  10  3


0.4 wt% 10 0.31 0.95 2.6  10  3 5 9.6  10–14 54 23 4  10  3
1.3 wt% 10 0.31 0.46 0.72 5 9.6  10–14 43 35 0.76
2.1 wt% 10 0.31 0.19 5.93 5 9.6  10–14 39 54 3.97
4.3 wt% 10 0.31 0.09 49 5 9.6  10–14 28 77 57

contact angle even surpasses the 901, meaning a shift from


hydrophilic to hydrophobic behavior.
In Fig. 6, the capillary number is represented graphically. We
have plotted both the capillary number calculated for the actual
evaporation rate and the evaporation rate of water without pre-
sence of MHEC. One can clearly see that for low MHEC concentra-
tions, the capillary number is less than 1, leading to a homogenous
drying behavior as observed in Fig. 4(a, b). This means that capillary
pressure differences are sufficient for driving the flow to the surface.
Note that the drying behavior changes at Ca1. The system
switches its behavior when 1.5 wt% MHEC (which coincides with
the concentration approximately used in practical applications) is
present in the pore solution. At the corresponding capillary number
of Ca 1, a front of the size of the sample should be visible.
At higher concentrations (2.1 and 4.3 wt% MHEC) Ca»1, corresponds
to a sharp drying front, meaning that the viscous force dominates
over capillary force and capillary force is no longer sufficient to Fig. 6. Capillary number as a function of MHEC concentration.
cause flow. One can observe a change of capillary number of over
four orders of magnitude, which is comparable to the change
observed in viscosity. The changes in the cosine of the contact
angle and/or surface tension only influence the capillary number of
an order in magnitude (in the same direction as the viscosity). The
decrease of surface tension and increase in contact angle will
reduce the capillary pressure at high viscosities and will as such
reduce capillary pressure, enabling the viscous forces to overcome
the capillary forces earlier.

4.2.2. Moisture diffusivity


To establish what is the dominating parameter responsible for
the change in the drying behavior, we will determine the moisture
diffusivity as a function of MHEC concentration from the measured
NMR profiles. In Fig. 7, the moisture diffusivity is calculated from
the NMR profiles of all different MHEC concentrations. One can
clearly observe that with increasing concentration of MHEC, the
moisture diffusivity at high moisture content decreases, indicated
by the arrow. Furthermore, we observe that at high moisture Fig. 7. Moisture diffusivity determined from the moisture profiles as a function of
contents, the minimum moisture diffusivity is constant over a MHEC concentration. The lines represent fits of the data.
large range of moisture contents (Θ). At these high moisture
contents, the flux is mainly dominated by liquid flow induced by
capillary pressure differences inside the porous material. In the
theory section, we have shown how flow is incorporated in the the moisture diffusivity for all viscosities simultaneously at high
moisture diffusivity. The flow scales with the inverse of viscosity moisture concentrations, two terms (first two on the right hand side
(q ¼  ðK=ηÞ∇p). Therefore, the moisture diffusivity should scale of the equation) instead of one are needed. This gives the following
with viscosity at high moisture content. new empirical function:
In order to understand the drying process more accurately, it may  1    1  1
D0l eβl Θ þ D1l eβl Θ þ Dmin þ Dv eβv Θ
0 1
be characterized by an empirical moisture diffusivity equation with a D¼ ð13Þ
limited amount of parameters as proposed by Landman et al. (2001)
(Landman et al., 2001). The physical meaning of this equation will be
Note that as discussed above, the constant ðD0l ) referring to the
discussed at a later stage after the fitting process. To accurately fit our
liquid flow should scale in with viscosity, which is in this equation
data for the moisture diffusivity, we use a similar empirical equation
represented by
with two changes. Based on the calculated moisture diffusivity, it is
required to define an explicit minimum. This is achieved by adding η0
D0l ðηÞ ¼ D ; ð14Þ
the moisture diffusivity Dmin to the equation. Secondly, in order to fit η η0
626 A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628

where η0 is the viscosity of water, η the visicosity of the pore fluid cellulose ether solution remained constant. Nevertheless, a reduc-
depending on the MHEC concentration and Dη0 the diffusion tion of evaporation rate was observed. This indicates that another
constant for water. From Eq. (14), moisture profiles are calculated mechanism is at play. We suggest that cellulose may be transported
using software package Comsol Multiphysics, using a coefficient to the surface and forms a skin. This hypothesis is supported by
form PDE. To obtain the most accurate fit, the moisture diffusivity Fig. 1c, which shows that in that event MHEC layer is formed, it will
as a function of moisture content was fitted simultaneously with be hydrophobic and as such free of liquid, thereby possibly creating
the calculated profiles as a function of position and time. The a barrier for vapor transport. 2) In Fig. 8, the values for Dmin are
results of these fits are shown by continuous lines in Fig. 5, Fig. 7 given, which were optimized to obtain the most suitable fit in the
and Fig. 9 representing respectively the total moisture content, profiles and integrals. The reason for the observed trend for Dmin is
moisture diffusivity and profiles. unclear. Several explanations are possible. First of all, both the
In Table 2, the fitting parameters are given for the empirical evaporation induced moisture velocity and Dmin show an apparent
moisture diffusivity equation, except for Dmin and the evaporation correlation (Fig. 8). This would support a hypothesis that Dmin is in
rate, which depend on the MHEC concentration. For the evapora- fact a vapor dominated diffusion flux. In this case, the flux due to
tion, this is apparent from the total moisture content as given in the evaporation process should be equal to the vapor diffusion. This
Fig. 5, which decreases with increasing MHEC concentration. The hypothesis remains to be proven. 3) The profiles and moisture
fitting process determined typical parameters for the surface layer diffusivity can be calculated based on scaling the diffusivity at high
(Dair =δ). The dependence of Dmin on MHEC concentration can be moisture content with one single other key parameter being
observed in Fig. 7. For both Dmin and v, the best values are plotted viscosity. Consequently, we conclude that the viscosity of the fluid
in Fig. 8. Note that the plotted values for v (out flux of moisture) is key for the change observed in the moisture diffusivity at high
are almost equal to those determined by the initial slope of the moisture content.
total moisture content plots ( Fig. 5). The higher error is due to the
fact that the barrier properties can vary as a function of time,
which is based on the barrier properties and Kelvin equation. We 5. Conclusions
observed that Dmin varies slightly as a function of concentration. At
later stages in the drying process, Dmin plays a dominant role. This NMR drying experiments performed on fired clay brick saturated
constant determines the decrease of moisture content in case of a with pure water show a well-known drying behavior. Until a critical
front. At high MHEC concentrations, the moisture diffusivity saturation is reached, the water is distributed homogenously. After
remains constant over a large range of the moisture content, in this critical saturation is reached, a drying front develops. In case of
that range the capillary forces are insufficient to drive the flow. drying of a MHEC saturated fired clay brick (FCB), inhomogeneous
The fitting process and presented observation leads to three key drying is observed, which results in a receding drying front at the
conclusions. 1) No difference in evaporation rate was expected, with start of the drying process. The capillary number confirms and
or without presence of MHEC, since our experiments of capillary explains this shift from homogenous to front receding drying. At a
tubes or bulk solutions have shown that evaporation rates of capillary number of around 1, the shift from homogeneous drying to
front receding drying occurs. This happens around 1.5 wt% of MHEC
Table 2 in pore solution, which approximately coincides with the concen-
Fitting parameters of the moisture diffusivity tration used in glue mortars. At higher concentrations (41.5 wt%
function. MHEC), Ca»1 and a front receding drying is observed; at low
concentrations (o1.5 wt% MHEC), Ca«1 homogeneous drying is
Parameters Value
observed.
Dη0 8.0  10  9 Analysis of the capillary number shows, that the main parameter
β0l 28 determining this change are the evaporation, viscosity, contact
D1l 1.0  10–11 angle and surface tension. The other parameters, such as porosity,
β1l 140 permeability, pore radius and sample size are constant throughout
Dv 10  7 the experiments, since they are properties of the material. The
βv 103 parameter showing the largest change is the viscosity, which
indicates that this is probably the key parameter influencing the
drying.
The moisture diffusivity was determined from the NMR profiles
by modifying the empirical moisture diffusivity equation of Land-
man et al. (2001) (Landman et al., 2001). Viscosity was the only
parameter changed to fit all five moisture diffusivity curves and
NMR profiles having the same set of parameters, combined with a
variation of the evaporation and a minimum moisture diffusivity
as a function of MHEC concentration. From the final values, we
observed that 1) the evaporation rate drops with increase MHEC
concentration, 2) the diffusivity scales with viscosity and 3) and a
clear minimum diffusivity is found.
The fact that the evaporation rate depends on the MHEC
concentration is quite unexpected, as experiments on MHEC solu-
tions in capillaries have all shown that evaporation is independent
of the MHEC concentration. In contrast, in an inert porous medium
such as a brick, it appears that the evaporation rate is reduced
immediately from the start of the drying process. It may be
Fig. 8. Minimum moisture diffusivity determined as a function of MHEC concen-
suggested cellulose is transported to the surface and forms a skin.
tration, including the fluid velocity induced by the evaporation at the top surface of This hypothesis is supported by Fig. 1c, which shows that in that
the sample. event MHEC layer is formed, it will be hydrophobic and as such free
A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628 627

Fig. 9. Moisture profiles as measured by NMR (indicated by the symbols) and calculated profiles (indicated by the lines) for different concentrations of MHEC a) 0 wt%
b) 1.3 wt% and c) 4.3 wt%. For 0 wt% the first 7 profile are given every 20 min and afterward after 40 min. For the 1.3 and 4.3 wt% the between profile is 60 min.

of liquid, thereby possibly creating a barrier for vapor transport. diffusion should be equal. This hypothesis remains to be proven.
This hypothesis remains to be proven. This observation has become apparent especially in this study since
We conclude that viscosity is indeed the main factor influen- to our knowledge previously no measurements were performed on
cing the type of drying behavior of the MHEC filled brick. The drying of high viscous fluids inside porous materials.
small change in contact angle and/or surface tension as a function
of MHEC does not influence the behavior significantly, as shown
by the high quality of the fits which are accurate by only scaling Acknowledgments
with viscosity. The fast change of regime from capillary flow to
front receding drying can be explained by the following: In case of This research was carried out under the project number
front receding drying, contact angle and surface tension are not M81.6.08315 in the framework of the Research Program of the
relevant, since the moisture is stationary and contact angle and Materials innovation institute (M2i) (www.m2i.nl). The authors
surface tension do not change this. Large changes in contact angle thank C.J. Haecker from SE Tylose GmbH & co, Wiesbaden,
and surface tension occur at high MHEC concentrations. At these Germany, for providing the cellulose ether samples and for helpful
high MHEC concentrations, the fluid has a high viscosity, which discussion in an early phase of the project. The authors would also
does not allow the capillary force to overcome the viscous forces. like to thank Hans Dalderop and Jef Noijen for their technical
In addition, the change of contact angle and surface tension only assistance.
decreases the capillary forces, allowing viscous forces to dominate
at lower concentration. As such it is conceivable that in a transi- References
tional regime (Ca 1), contact angle and surface tension may
reduce the concentration at which the capillary number changes Chauvet, F, Duru, P, Geoffroy,, S, Prat, M, 2009. Three periods of drying of a single
from below 1 to above 1. square capillary tube. Phys. Rev. Lett. 103, 124502.
Genuchten, V., 1980. A closed-form equation to predicting the hydraulic conduc-
The fact that the moisture diffusivity Dmin remains constant over
tivity of unsaturated soil, M. Th. Soil Sci. Soc. Am. J 44, 892–898.
a large range of moisture contents at high MHEC concentration and Hall, C, Hoff, D, 2012. Water Transport in Brick, Stone and Concrete,. Spon Press,
the apparent relationship with the moisture velocity at the surface London.
of the sample, support a hypothesis that Dmin is in fact a vapor Kopinga, K., Pel, L., 1994. One-dimensional scanning of moisture in porous
materials with NMR. Rev. Sci. Instrum 65, 3673–3681.
dominated diffusion flux. In case of a constant gradient in moisture Landman, K.A., Pel, L., Kaasschieter, E.F., 2001. Analytic modelling of drying of
content, the evaporation process and the corresponding vapor porous materials. Math. Eng. Ind 8, 89–122.
628 A.P.A. Faiyas et al. / Chemical Engineering Science 123 (2015) 620–628

Le Bray, Y., Prat, M., 1999. Three-dimensional pore network simulation of drying in Prat, M., 1993. Percolation model of drying under isothermal conditions in porous
capillary porous media. Int. J. Heat Mass Tran. 42, 4207–4224. media. Int. J. Multiphas. Flow 19, 691–704.
Lenormand., R, Touboul., E, Zarcone., C, 1988. Numerical models and experiments Prat, M., 2002. Recent advances in pore-scale models for drying of porous media.
on immiscible displacements in porous media. J. Fluid Mech. 189, 165–187. Chem. Eng. J. 86, 153–164.
Lu, T.X., Biggar, J.W., Nielsen, D.R., 1994. Water movement in glass bead porous Prat, M, 2007. On the influence of pore shape, contact angle and film flows on
media: 1. Experiments of capillary rise and hysteresis. Water Resour. Res 30, drying of capillary porous media. Int. J. Heat Mass Tran. 50, 1455–1468.
3275–3281. Prat, M, 2011. Pore network models of drying, contact angle, and film flows. Chem.
Metzger, T., Tsotsas, E, 2005. Influence of pore size distribution on drying kinetics: a Eng. Technol. 34, 1029–1038.
simple capillary model. Dry. Technol. 23, 1797–1809. Prat, M, Veran-Tissoires, S, Vorhauer, N, Metzger, T, Tsotsas, E., 2012. Fractal phase
Metzger, T., Tsotsas, E, 2008. Viscous stabilization of drying front: three-dimensional distribution and drying: impact on two-phase zone scaling and drying time
pore network simulations. Chem. Eng. Res. Des. 86, 739–744. scale dependence. Dry. Technol. 30, 1129–1135.
Pel, L., Kopinga, K., Brocken, H., 1996. Moisture transport in porous building Shokri, Nima, Sahimi, Muhammad, 2012. Structure of drying fronts in three-
materials. Heron 41, 95–105. dimensional porous media. Phys. Rev. E 85, 066312.
Pel, L., Brocken, H., Kopinga, K., 1996. Determiination of moisture diffusivity in Yiotis, A.G., Stubos, A.K., Boudouvis, A.G., Tsimpanogiannis, I.N., Yortsos, Y.C., 2005.
porous media using moisture concentration profiles. Int. J. Heat Mass Tran. 39, Pore-network modeling of isothermal drying in porous media. Transport
1273–1280. Porous Med. 58, 63–86.

You might also like