You are on page 1of 41

Chapter Four

Results and Discussion

4.1 Introduction

This chapter represent the result of preparation parameters on synthesized


process of prepared adsorbent (Ca/Fe)-LDH and immobilized it as sodium alginate
beads to used it as a reactive material to reduce the concentration of selected (TC
and AMX) antibiotics from contaminated solution. Also discuss the results of
characterization tests to describe the composition, functional group and
morphology of the synthesized adsorbent by applying Field Emission Scanning
Electron Microscope (FESEM), N2 adsorption/ desorption (BET) technique and the
X-ray diffraction (XRD) analysis.

In addition, this chapter also deals with finding the optimum values of
different operational conditions such as contact time , initial solution pH ,
influence of agitation speed, adsorbent alginate beads mass and initial antibiotics
concentration based on maximum antibiotics concentration removal efficiencies
and discuss the results of batch study to reach to how the parameters effect on the
ability of (Ca/Fe)-LDH-sodium alginate beads for the sorption of selected
antibiotics (TC and AMX) from contaminant water by application laboratory
batch. Isotherm, kinetic and thermodynamic analysis are also studied.

The adsorption mechanism of TC and AMX antibiotics from contaminated


solution on the synthesized (Ca/Fe)-LDH-sodium alginate beads was analyzed and
illustrated by using characterization tests Fourier Transform Infrared Spectrometer
(FT-IR) with Energy-dispersive Spectrometer EDS) before and after adsorption
process to provide a theoretical basis for the use of adsorbent beads in antibiotics
removal processes and highlight its practical applicability.
Continuous experimental system used fixed packed bed model composed of
single phase fixed bed within a column to simulate the actual operation of barrier
for the antibiotic removal using (Ca/Fe)-LDH-sodium alginate beads. To
understanding of the dynamic behavior of adsorption systems which helps in
finding the design parameters of this column, this work used different operating
parameters (initial adsorbate concentration, bed height and flow rate) for
evaluating the efficiency of adsorbent in a continuous treatment process and effect
on breakthrough curves. Finally, all column experiments collected data have been
modeled by five mathematical models (Adams-Bohart Model, Thomas-BDST
Model, Yan Model, Belter-Cussler-Hu Model and Clark Model) which used to fit
the experimental results

4.2 Synthesis of (Ca/Fe)-LDH-Sodium Alginate Beads

The extraction of calcium ions from chicken eggshells is the initial step in
the synthesis of (Ca/Fe)-LDH nanoparticles prior to their entrapment in the
polymer matrices of sodium alginate, and it calls for the mixing of one gram from
eggshell powder.(after collect from the source, rinse several time with water then
dried, crushed, dried at 105 ˚C and passed on sieve to obtain 0.6 mm) with 50 mL
solution consisted of distilled water and specific amount from 35-38%
hydrochloric acid

By adding 5, 10, 15, and 20 milliliters of hydrochloric acid to 100 milliliters


of distilled water containing 1 gram of chicken eggshells, we were able to identify
the optimal amount of HCl needed to extract the maximum amounts of calcium
ions from the shells. The solutions are stored in four beakers and stirred vigorously
for three hours using a magnetic stirrer..
After each solution has been filtered through paper, its calcium ion
concentration can be measured by atomic absorption spectroscopy. As shown in
figure 4.1 , increasing the volume of HCl used to extract calcium ions from
chicken eggshells resulted in a change from 199000 mg/L to 200389 mg/L in
concentration of these ions. Significantly more calcium was extracted as the
volume of acid was increased from 5 to 10 mL, reaching a concentration of 200000
mg/L; after this amount, only a little rise in the number of Ca+2 ions was seen.
Because of the excess free solution they leave behind, these ions can cause a
complicated component that may interact with the carbonation reaction.(Kim &
Jung, 2020). The aqueous solution containing the highest concentration of Ca+2
extracted from eggshells was thus prepared using a volume of HCl of 10 mL

Figure (4.1): Calcium ion concentrations extraction from egg shell with
various amounts of HCl

The effect of three operational parameters on the production of calcium iron


lanthanum disulfide-sodium alginate was studied. These were the solution's pH, the
calcium-to-iron molar ratio, and the amount of (Ca/Fe)-LDH nanoparticles added
to the alginate solution. Maximum efficiency in removing TC and AMX from
polluted water under varying these parameters are a useful metric for determining
optimal values..

By varying the water's pH from 7 to 12, with all other parameters held
constant, we were able to investigate how changing the water's pH affected the
sorption efficiency of the produced nanoparticles sorbent. The initial contaminated
solution pH was 7 and the agitation speed was 200 rpm for 3 hours, and alginate
mass was 0.5 g/100 mL for both TC and AMX. Sodium alginate's effectiveness in
the elimination of the antibiotic under consideration is shown to be pH-dependent
in Fig. (4.2).

Figure (4.2): Remediation of TC- and AMX-tainted water using (Ca/Fe)-


LDH-sodium alginate beads: effects of starting pH
Maximum efficiency was observed at a pH of 12, where values of 40.2% for
TC and 32.2% for AMX were recorded; however, efficiency was reduced to 34.2%
and 27.8%, respectively, when the pH was lowered to 7. This may have been
caused by an increase in the diameter of the nanoparticles.(Saha et al., 2018).

In the end, (Ca/Fe)-LDH-Na-alginate, which would be employed in the


subsequent tests, was manufactured at a pH of 12. At pH 12 and a dosage of
(Ca/Fe)-LDH nanoparticles to alginate solution equal to 5 g/100 mL, a series of
studies were conducted to assess the impact on bead production of a (Ca/Fe) ratio
spanning (0.5 - 4). In Fig.(4.3), we can see that the highest removal efficiencies for
TC and AMX antibiotics can occur at molar ratio 1 when sorption experiments are
performed at the same circumstances selected for starting pH.. Changes in the
radius difference between iron and calcium atoms or disorder in the structure of the
adsorbent used in LDHs could account for the lower efficiency compared to the
best possible value (Milagres et al., 2017).

The final step was to analyze how different concentrations of (Ca/Fe)-LDH


nanoparticles impacted alginate bead formation. In any case, one can examine
sorption efficiencies for this parameter in Fig. 1., which shows that the mass of
nanoparticles added to the alginate solution to immobilize it and make beads
changed from 1 to 6 g/100 mL at best values that obtained previously of initial pH
and molar ratios of (Ca/Fe) specified from previous tests (4.4).This number
indicates that removal efficiency of adopted antibiotics dramatically increased with
increasing nanoparticles dosage from 1 to 6 g, and that adding 0.5 g (Ca/Fe)-LDH
alginate beads per 50 mL of contaminated water may remove 41.2% of TC and
32.3% of AMX antibiotics, respectively..

Figure (4.3): Remediation of TC- and AMX-contaminated water using


(Ca/Fe)-LDH-sodium alginate beads: effects of (Ca/Fe) molar ratio

Figure (4.4): Dosage effect of Ca/Fe)-LDH nanoparticles on Na-alginate for


making adsorbent beads for treating water tainted with TC and AMX.

4.3. Characterization of (Ca/Fe)-LDH-sodium alginate

4.3.1 Morphology of synthesis adsorbent

SEM graphs (4.5) and (4.5) provide details about the morphological features
of (Ca/Fe)-LDH-sodium alginate before and after interaction with TC and AMX
antibiotics (4.6). The nanoparticles had more surface area as a powder than as
beads and adsorption capacity than the beads, but the surfaces appear compact and
disordered compared to before immobilization with sodium alginate, as seen in this
figure, which was taken at magnifications of 200 and 500 nm.(Vakili et al., 2020).
Particle size distribution with mean size of 4.5 nm; yet, this surface can be
highly disordered and compact. The beads appear to have a macro-porous surface
that facilitates the adsorption of oxyanions, and they also have cavities with
protruding and receding surfaces, both of which contribute to an increase in
antibiotic adsorption.(Fan et. al., 2022) and have been more exploited by TC than
AMX, as illustrated by laboratory experiments, which first show that TC has a
greater affinity for removal than AMX.
Compared to the smooth surface of sodium alginate without nanoparticles,
these beads' (consisting of Ca and Fe(III)) surface is broken and uneven, reflecting
the weaker mechanical resistance. Due to the aggregation and coating of the
sorbent surface by TC and AMX antibiotics molecules, the morphology of the
adsorbent beads is significantly different after the sorption process than it was
before.(Mansour et. al., 2022) and this clearly shown in figure (4.6) as schematic
diagram of SEM images with 100µm magnification scale.

Using the N2-adsorptione/desorption method of BET analysis, we were able


to calculate the specific surface area and total pore volume of the synthesized
adsorbent material, and we found that the values for (Ca/Fe)-LDH and (Ca/Fe)-
LDH-sodium alginate bead were, respectively, 19.52 m2/g and 15.21 cm3/g. The
average pore diameter of (Ca/Fe)-LDH-sodium alginate and (Ca/Fe)-LDH-sodium
alginate beads was estimated to be (0.0178, 0.0163 ) nm using the BJH method.
Adsorbent beads are able to allow molecules of antibiotics to enter, interact with
surface functional groups, and increase adsorption because they have an adequate
surface area and total pore volume..
The microstructure of the (Ca/Fe)-LDH-sodium alginate beads was studied
using energy dispersive spectroscopy (EDS). EDS graphs, such as the one shown
in Fig. (4.7) and reported in Table, can be used to estimate the relative abundances
of individual elements in (Ca/Fe)-LDH-sodium alginate and to specify the change
and the increase magnitude that happened in these relative abundances beyond the
sorption process (4.1).
The elemental analysis of the adsorbent beads is summarized in the
following table. C, Na, N, O, S, Fe, Ca, and H were found to account for 14.21%,
3.81%, 0.60%, 0.05%, 11.31%, 5.61%, and 5.01% of the total dry weight of the
prepared sorbent, respectively; this, in turn, explains the existence of a striking
increase in the C and N content beyond the sorption of TC, to 15.21% and 6.11%,
respectively, which certifies the adsorption of TC (4.7b). Similar trends were seen
in the EDS examination of the adsorbent beads after they had been loaded with
AMX, with the percentages of C, N, and S elements increasing to 15.01 percent,
5.45 percent, and 1.01 percent, respectively (Fig (4.7 C). These percentage changes
indicate that antibiotics and prepared (Ca/Fe)-LDH-sodium alginate beads interact
successfully.

Figure (4.5): (SEM) graphs for a) (Ca/Fe)-LDH as nanoparticles, b) (Ca/Fe)-


LDH-sodium alginate beads before, c) and d) after interaction with TC and
AMX antibiotics at magnification scale of 200 and 500 nm.

Figure (4.6): Schematic diagram of SEM images illustrate surface morphology change for
(Ca/Fe)-LDH surface before and after immobilization to beads as well as (Ca/Fe)-LDH-Na
alginate beads after interaction with TC and AMX for magnification scale = 100µm
Figure (4.7): Graphs of EDS for a). (Ca/Fe)-LDH-sodium alginate beads
before interaction with antibiotics and b). and c). after sorption of TC and
AMX antibiotics.

Table (4.1): Energy-dispersive X-ray spectroscopy (EDS) examination of


(Ca/Fe)-LDH-sodium alginate both before and after sorption of TC and AMX
antibiotics revealed the following elemental percentages:.

4.3.2 X-ray diffraction (XRD) analysis

Sodium alginate (used for immobilization of (Ca/Fe)-LDH) was synthesized


in the experimental approach described above, and its crystalline composition was
determined by X-ray diffraction (XRD) analysis, as shown in Fig (4.8). It is
necessary to scan the sample through a range of 2 angles (from 20 to 80) to get all
possible lattice diffraction directions, and numerous diffraction reflections can be
recognized at specific intensities (13.5, 20.4, 32.2, 43.6, and 45.1). Antibiotics (TC
and AMX) are removed from aqueous solution via active sites on the surface of
sodium alginate, which are represented by these reflections.. In comparison with
“Joint Committee on Powder Diffraction Standards (JCPDSs)”, these reflections
have been identical to the (Ca/Fe)-LDH nanoparticles, calcium carbonate (CaCO 3)
calcium hydroxide (Ca(OH)2) and iron nitride )Fe2N( as evidence from Fig. (4.8).
the presence of (Ca/Fe)-LDH nanoparticles in the immobilized sodium alginate
refer to the success of formation of such component (Ruan et al., 2011; Szabados
et al., 2018).
Figure (4.8): X-ray diffraction (XRD) pattern revealing the crystalline
structure of (Ca/Fe)-LDH-sodium alginate.

4.3.3 The FTIR spectrum of the compound (Ca/Fe)-LDH-sodium alginate

One of the best techniques for characterizing layered double hydroxides (LDH)
is Fourier transform infrared spectroscopy (FTIR), which measures infrared
intensity vs. wavelength of light over a range of 400 to 4000 cm -1. FTIR is used to
identify the functional groups that are present on the surface of nanoparticles when
hydroxyl, carbonyl, and other groups are present, as these groups function as a
source for the synthesis of adsorbent (Reddy et al.,.2022).
Figure (4.9) displays infrared absorption spectra for (Ca/Fe)-LDH as
nanoparticles and (Ca/Fe)-LDH as sodium alginate beads before and after
antibiotics sorption to identify the key functional groups that enhanced the
contaminant adsorption, as the adsorption capacity of synthetic adsorbents depends
on chemical surface function groups in addition to surface area.(Palacio et. al.,
2022).
Intense absorption bands were found to form as a result of hydroxyl (-OH)
group stretching vibrations and amide stretching vibrations (N-H). The broad and
strong absorption band at frequency (3570-3000 cm-1) observed for (Ca/Fe)-LDH-
sodium alginate was previously observed in nanoparticles and had shifted
downward and outward due to the stretching mode of OH-, (N-H) groups,
stretching vibration of the hydrogen bond, or the formation of interlayer water
molecules..(Chen et al. 2018).
Spectra of (Ca/Fe)-LDH-sodium alginate showed two absorption bands, one
with moderate intensity at 1597 cm-1 due to asymmetric and symmetric stretching
vibrations of the carboxylate group, and the other with weak intensity at 1411 cm-1
(-COO )( Daemi & Barikani 2012 ; Aprilliza 2017) Moreover, the stretching
vibrations of the (-C-O) and (-C-C-) bonds can be seen as two weak absorption
bands in the spectrum, located at 1126 cm-1 and 1026 cm-1. The (-C-H) bond's
asymmetric and symmetric stretching vibrations were less prominent in both
representations, and they were obscured by the hydroxyl group's strong, broad
stretching vibrations.(Badita et al., 2020).
From Fig. (4.9) the band broadening is observed by intercalation that results from
the electrostatic interaction between TC molecules and hydroxide layers to suggest
their safe stabilization in the interlayer space of LDH (Mallakpour & Dinari,
2015). Also it showed some peaks after TC adsorption disappeared, shifting or
decrease in its intensity, the peaks that disappeared after TC adsorption showing
active involvement of these peaks in the adsorption process (Khan et al. 2012). The
hydroxyl (-OH) and amide (N-H) group stretching vibrations were responsible for
the broad, intense absorption band at frequency (3317 cm-1), Moreover, the
medium and mild absorption bands at 1627 cm-1 and 1415 cm-1, respectively, can
be attributed to the asymmetric and symmetric stretching vibrations of the
carboxylate group (-COO-). Also, the stretching vibrations of the (-C-O) and (-C-
C-) bonds showed up as two weak absorption bands in the spectra, at the two
frequencies (1087 cm-1 and 1029 cm-1)
Also from fig(4.9) we can noticed that the change in peaks after AMX
adsorption with stretching vibration of hydroxyl (-OH) and amides (N-H)groups at
frequency (3255 cm-1) and asymmetric and symmetric stretching vibrations of the
carboxylate group (-COO-) appeared at the frequency (1631 cm -1) the frequency
(1415 cm-1)respectively. The stretching vibrations of the (-C-O) and (-C-C-) bonds
also displayed at the two frequencies (1087 cm-1, and 1029 cm-1) respectively.
While the asymmetrical and symmetrical stretching vibrations of the (-C-H) bond
did not appear because it overlapped with the strong and broad stretching
vibrations of the hydroxyl group. Also noticed that the Sulfate and sulfoxide
groups at removal of AMX due to S=0 stretching vibration shifted at frequencies
(1415 cm-1 to 1380 cm-1 ) and (1070 cm-1 to 1030 cm-1) (Ragab et. al., 2021).
Accordingly, successful removal of adopted antibiotic occurs on the adsorbent
surface with supporting of mentioned groups.

Figure (4.9): Before and after sorption of TC and AMX antibiotics, the FT-IR
spectrum of (Ca/Fe)-LDH nanoparticles and (Ca/Fe) LDH-sodium alginate
was examined.

4.4 Operation parameters in batch testing

4.4.1 Contact Time

The contact time of antibiotics adsorption over the adsorbent is one of the
important factors affecting the process kinetic and specify the time required for
contaminant transmission from solution and diffusion on the adsorbent surface
(Nguyen et al., 2022).
The time required to reach the equilibrium is achieved by batch study through
the antibiotic concentrations change with the time. Fig. (4.10) shows how
operating conditions of Co 100 mg/L, pH 7, beads mass 0.5 g/100 mL, and
agitation speed 200 rpm affect the efficiency with which TC and AMX are
absorbed from contaminated water onto produced (Ca/Fe)-LDH sodium alginate
beads for contact times not exceeding three hours. Antibiotics at a concentration of
100 mg/L are eliminated fast until equilibrium is reached or total absorption
occurs, resulting in a decrease of 39.4 percent TC and 30.3 percent AMX after 1.5
hours (90 min).
Antibiotics are eliminated at a high rate initially, but this rate slows after 90
minutes as the number of binding sites for sorption of antibiotic molecules
decreases.(Faisal & Naji, 2019), Mean equilibrium time for TC and AMX
adsorption was obtained in under 1.5 hours, although there was no noticeable
change in removal percentages until 180 minutes..
The increase in the adsorption time does not make the adsorption rate obviously
change since the removal percentages had been slowly tended to be stable till 180
min because of the quantity of free sites on adsorbent surface were reduced which
cause the binding of adsorbate from the aqueous solution will be harder and the
sorption processes thus decreases (Lucaci et al., 2021).

Figure (4.10): Influences of time on the removal of TC and AMX onto


adsorbent beads

4.4.2 Initial Solution pH

One of the most important elements in the removal of antibiotics is the pH of


the contaminated water, which affects the surface charges of the beads, the degree
of TC and AMX ionization, and the dissociation of functional groups.(Luo et al.,
2020). Fig. (4.11 ) Antimicrobial removal efficiency versus baseline pH values
ranging from 3 to 12 as a scatter plot. Dimethyl ammonium, tri-carbonyl amide,
and phenolic diketone groups are only a few of the ionizable functional groups
found in the TC molecule.(Hsu et al., 2018).
With pKa values of 3.3, 7.7, and 9.7, TC molecule may exist as the cation
(TC+00), zwitterion (TC+-0), monoanion (TC+--) and dianion (TC0--) form,
respectively (Qin et al., 2018), therefore; the removal ratios were low (≤17.5% and
≤14.9% at pH 3.0) for TC and AMX respectively. It's possible that antibiotic-H+
ion competition is at the root of this behavior..
In the case of AMO adsorption onto adsorbent beads with pKa values of 2.67,
7.11, and 9.55, the same outcomes were observed. When the pH is raised closer to
neutral, the sorption efficiency increases dramatically, reaching not less than
40.2% for TC and 32.1% for AMX antibiotics at pH 7. When an antibiotic is in a
neutral environment, its ionization and hydration levels drop, facilitating
elimination via hydrogen bonds and π– π stacking effect.
The same chart also reveals that a return to the solution's fundamental form is
associated with a drop in removal efficiency, while efficiencies for both TC and
AMX are above 25.6% and 18.2%, respectively. The weakening of hydrogen
bonds, which may have resulted from the production of OH- at higher pH levels,
may explain the efficiency decreases observed at these conditions.. From figure
(4.12a) the pHpzc value of (Ca/Fe)-LDH nanoparticles was determined to be 7 and
for alginate beads measured as around 7.18.
Fig(4.12b) denotes that at pH < pH pzc ,the adsorbent surface is positively
charged due to ions of H+ are transferred to the particle surface leading to a
positive adsorbent beads surface and combined with OH- groups of antibiotics by
the electrostatic attraction forces that generated between different charges and this
will enhance the sorption of antibiotics and this illustrated as the initial pH went
up, equal a decrease of zeta potential from range (2-6) so there was increase in the
adsorption of antibiotics onto (Ca/Fe)-LDH-Na-alginate beads and that is agreed
well with the results above(Álvarez-Esmorís, et al., 2022).
The levels of TC and AMX bound to the adsorbent, which served as adsorption
sites for the antibiotics, were, on the other hand, not leaked out at higher pH values
(pH > 6). As a result, the adsorption capacity of TC and AMX increased as the
interaction between antibiotics and beads increased. When the pH was higher than
the pKa value, the adsorption capacity dropped because the adsorbent surface
became negatively charged after combining with the H+ groups of the antibiotics,
resulting in electrostatic repulsion between the charge of the beads' surface and the
OH- groups of the antibiotics. Other experiments for TC and AMX adsorption by
this material were performed at pH 7. Competition between antibiotics and surface
functional groups such as -COOH and -OH can also contribute to a decrease in the
sorption at higher pH..

Figure (4.11): ) Initial pH effect with removal of TC AND AMX from


contaminated water

Figure (4.12): (a) pH PZC for (Ca/Fe)-LDH nanoparticles , (b) pH PZC of


(Ca/Fe)-LDH-sodium alginate beads.

4.4.3 Influence of Agitation Speed


The adsorption experiments were carried out as a function of agitation by the
variation of concentration with different agitation rates .The agitation percentage
influenced highly that of adsorption which rises when the speed of agitation rises.
According to figure (4.13), the increase in agitation speed is essential to
enhance TC and AMX removal from polluted water because of diffusion rate of
antibiotics molecules rise towards the adsorbent beads surface under operational
parameters of time=90 min, pH=7, 0.5 g/100 mL beads mass, and primary
adsorbate =100 mg/L concentration of.
Maximum shaking speeds allow antibiotic molecules to come into touch with
alginate beads, which improves their transit to the solid phase and hence their
removal effectiveness.(Asmaly et al., 2015). Removing TC and AMX efficiencies
was not bigger than 11.2% and 9.4% for static aqueous solutions, although
dramatically rose at 40.2% and 32.1% respectively at 200 rpm. Yet, the figure also
explains that the adsorption rises when agitation speeds rises only up to a specific
limit, so speed to 250 rpm increase does not make any changes in the removal
efficiencies because beyond there is no rise in the adsorption capacity because at
higher agitation speed it was attributed to an increased desorption tendency for
some loosely bound antibiotics molecules at the equilibrium time (Barquilha &
Braga, 2021).

Figure (4.13): Effects of agitation speed on the behavior of adsorbent beads


to remove antibiotics from contaminated water.

4.4.4 Adsorbent alginate beads mass (dosage) impacts


In specific, the impact of adsorbent dosage is the most significant factor to take
into consideration, since it defines the degree of adsorption and calculates the
expense of adsorbent each unit of solutions to process (Rathi & Kumar, 2021). The
percentage of removal increases with the increase of adsorbent dosage for the
initial concentrations studied, as shown in Fig. (4.14), which represents the impact
of the adsorbent alginate beads mass on the removal efficiency of TC and AMX
antibiotics. Clear evidence suggests that increasing the dosage of the utilized TC
and AMX antibiotics from 0.1 to 1.2 g can result in an increase in removal
efficiencies from the lowest value (13.3%) to become not less than (95%) and from
(9.8%) to more than (90%).
The larger the mass of the adsorbent alginate beads, the more adsorbent surface
area there is, which implies more empty sites are available to interact with the
molecules of pollutants, leading to greater removal efficiency.(Wei et al., 2020). It
is noteworthy that the removal percentages do not attain 100% with increasing the
adsorbent dosage up to 1.2 g (Figure 4.14) confirming the particle aggregation of
the adsorbent hindered complete removal of antibiotics (Takdastan, et al., 2016).
While, further adsorbent increase with more than the optimized value results in
a decrease in both the removal efficiency and the adsorption uptake. This decrease
may be due to a reduction in the total adsorption surface area available to
antibiotics, resulting from overlap or aggregation of adsorption sites which
consequently declines the vacant active sites (Sayğılı & Güzel , 2016).

Furthermore, this Agglomeration or aggregation of adsorbent mass effects on


increasing diffusional path length for antibiotics adsorption, lowering the
adsorption rate (Tony, 2022).

Figure (4.14): Relation between sorbent dosage and removals of TC and AMX
antibiotics

4.4.5 Initial Antibiotics Concentration

In order to examine the effect of the initial concentration of TC and AMX on


adsorption reactions, the antibiotics concentration was varied in the range of 50–
300 mg/L at 90 min, pH 7, mass of adsorbent beads =1 g/100 mL and 200 rpm for
agitation speed. Fig. (4.15) showed that at the established operational conditions,
the efficiency of TC antibiotic removal onto prepared alginate dropped
substantially from 77.4% at Co 50 mg/L to 20.5% at Co 300 mg/L, and for AMX
the efficiency dropped from 72.4% at Co 50 mg/L to 12.4% at Co 300 mg/L..
It is believed that all contaminating molecules are interacting with available
binding sites at lower concentrations, which undoubtedly may lead to a notable
improvement in the sorption effectiveness and accounts for the high removal rate
in the first stage of the adsorption process.(Abd Ali et al.2020).
As the contact time between the adsorbent and the adsorbate increases, the
active sites on the adsorbent become fully occupied by the adsorbed antibiotics
molecules, or the repulsive force between the antibiotic molecules on the surface of
the adsorbents beads and the antibiotic molecules in the bulk liquid phase
decreases the efficiency of the adsorption process.(Balarak et al., 2021). Although
adsorbent removed most of antibiotics at one and half hours under the same
conditions, possibly, the antibiotic showed higher removal if the concentration was
lower than the detecting limit as in environmental concentrations.

Figure (4.15): Effects of initial concentration on the adsorbent beads


behavior used to remove TC and AMX from contaminated water

4.5 Kinetics of Antibiotics Sorption

Adsorption kinetics from batch study are one of the most important features
of such processes that control the sorbent's capacity for adsorption and rate of
adsorption at the interface of solid-liquid phases due to the equilibrium state in
batch mode, the kinetics help researchers who conducted laboratory scale research
to estimate an adsorbent's performance before applied it on a larger scale (Debnath,
et al., 2020). Using (Ca/Fe)-LDH-sodium alginate beads, the sorption of TC and
AMX antibiotic over time was measured experimentally, and kinetic models were
developed for varying values of initial contaminant concentrations. For this
purpose, we used the "Solver" function in Excel 2016's non-linear regression
analysis toolkit to fit the measurements to pseudo first and second order models
(Eqs. 2.1 and 2.2), completing the formulation process.(Lagergren, 1898; Ho &
McKay, 1999).

Parameter values for fitted kinetic models are shown in Table (4.2);
nonetheless, this table provides independent confirmation that Model 2 is more
effective. Because the determination coefficient (R2) for TC is 0.99404 and the
sum of the squared errors (SSE) is 0.12148, and the determination coefficient (R2)
for AMX is 0.9924 and the SSE is 0.1287, we can conclude that the TC and AMX
interactions with the prepared adsorbent alginate beads are reasonably represented.
In addition, agreement between observations and kinetic models is evident in Fig
(4.16).

The sorbed amounts of TC and AMX onto alginate beads were measured to
be 4.4446 and 4.0288 mg/g, respectively, which are close to the calculated
amounts and are further evidence for the applicability of the second order model.
Thus, the chemi-sorption of the two antibiotics onto the alginate beads serves to
accomplish the sorption (i.e. chemical reaction).

Non-linear regression analysis was used to determine the values of the


constants in the model for intra-particle diffusion (Eq. 2.3) that relate qt to t 0.5.
(4.2)(Srivastava, et al.,1989). For each antibiotic tested, the relationship is depicted
by a straight line where the intercept does not equal zero, as shown in Fig. (4.17).
Therefore, this strongly implies that intra-particle diffusion is necessary for TC and
AMX sorption, but that the rate-controlling step will not be modeled.(Chen et al.,
2010).
Multi-linearity in the plots indicates the presence of multiple concomitant
pathways in the absorption of TC and AMX antibiotics. Lines in "part 1" appear to
have a steeper slope, suggesting that there may be more available sites for
adsorption of contaminants during the first stage of the sorption process. In
"portion 2," the rate-controlled step may have been intra-particle diffusion, which
is why the slope of the line is changed gradually. Part 3 denotes an equilibrium
stage that may be connected to diffusion slowing down because of a low
concentration of residual contamination in the liquid.(Cheung et al., 2007 ; Özer,
et al., 2009 ; Wang et al., 2010).

Table (4.2). Kinetic parameters for sorption of TC and AMX antibiotics onto
prepared adsorbent.

Figure (4.16): Kinetic models for interaction of (Ca/Fe)-LDH-sodium alginate


beads with with a). TC and b). AMX antibiotics solutions.

Figure (4.17): Intra-particle diffusion model for interaction of (Ca/Fe)-LDH-


sodium alginate beads with TC and AMX solutions.

4.6 Effect of Temperature and Thermodynamic Parameters

In order to investigate the effect of temperature on the adsorption of


antibiotics by adsorbent, the experiments were conducted at three temperatures
(22, 30 and 35 ˚C) as presented in Figure (4.18) with optimum operational
conditions at a concentration of 100 mg/L, pH 7, 90 min of contact time, and
agitation speed (200 rpm).The figure shows the change of removal efficiency of
each antibiotics with respect to varying temperature values. It can be noticed that
the removal efficiency reduced from 92.74% to 77.34% for TC and from 88% to
67% for AMX as the solution temperature increases from 22 to 35 ˚C.

. Figure (4.18): The effect of temperature on removal efficiencies TC and


AMX on (Ca/Fe)-LDH-sodium alginate beads at Co 100 mg/L, pH 7, 90 min
and 200 rpm.

This could be due to antibiotics molecular mobility have accelerated,


causing molecules to diffuse from the liquid phase to the solid phase more quickly
at high temperatures and this has led to a weaker contact between adsorbate and
adsorbent and a decrease in removal efficiencies. Furthermore, when the
temperature rises, adsorption removal decreases due to bond breaking on the
adsorbent surface, which weakens the attraction between active sites and
antibiotics, while at a lower temperatures, active sites expansion improved removal
efficiency and adsorption capacity, potentially increasing interaction with
antibiotics (Hamad & El-Sesy, 2023). Thermodynamic analyses were necessary in
investigating the behaviors of absorbent and mechanisms because its provide a
more descriptive interpretation of the adsorption process of TC and AMX
antibiotics on (Ca/Fe)-LDH-sodium alginate beads (Santos et al., 2020).

Thermodynamic Parameters such as ΔG°, ΔH°, ΔS° were calculated using


by a plot ln kd vs. 1/T at a temperature range (22-35 ˚C) were determined by using
equations (2.10-2.12). The values of ΔH° and ΔS° were obtained from the slope
and the intercept of the plotting as shown in figure (4.19) whereas Table (4.3)
presents thermodynamics data that obtained. The obtained values of the enthalpy
change ∆H° for both TC and AMX adsorption on (Ca/Fe)-LDH-sodium alginate
beads with (-68.522 kJ/mol) and (-61.75 kJ/mol) were in negative value indicated
that an exothermic nature for the adsorption process in the given conditions, so
increasing the temperature of the system does not usually favor the process and
leads to lower antibiotic uptakes. In addition, the magnitude of obtained values
which higher than 40 kJ/mol, provide information on the binding mechanisms
between antibiotics and the adsorbent which approve the chemisorption process
(Bansal, 2013; Li, et al., 2017).

The adsorption of antibiotic molecules on (Ca/Fe)-LDH-sodium alginate


beads was shown to be energetically stable by the negative values of entropy
change ΔS° for both TC and AMX, which indicated the decrease of randomness at
the solid/solution interface during the adsorption process. Furthermore, this value
may indicate a change in the structure of LDH due to the interaction through the
adsorption. A similar result was seen in (Santos et al., 2020; Yu et al., 2020).
The negative values of the Gibbs free energy change ΔG° confirm that the
adsorption of TC and AMX antibiotics on (Ca/Fe)-LDH-sodium alginate beads
was with a spontaneous character for all antibiotics tested. As temperature rises,
the values of ΔG° decrease (become more negative), indicating that the adsorption
process is more favorable at 22 °C (low temperature) (Pezoti et al., 2016; Li et al.,
2022). Table (4.3) reveals that as temperature increased, the magnitude of ΔG° for
adsorption reduced from (-8680.056, -6792.71) to (-6147.91, -4425.191) kJ/mol for
TC and AMX, respectively, indicating that the degree of spontaneity decreased.

Table (4.3). The thermodynamic parameters for sorption of TC and AMX


antibiotics onto prepared (Ca/Fe)-LDH-sodium alginate beads
Figure (4.19): plotting ln kd vs. 1/T for adsorption TC and AMX antibiotics
onto prepared (Ca/Fe)-LDH-sodium alginate beads

4.7 Sorption isotherms study

The isotherms are utilized to express the quantities of contaminant


molecules retained on the adsorbent beads and that remaining in the water at the
end of sorption process for equilibrium status. Maximum sorption capacity and
intensity distribution of contaminant molecules are the main characteristics of
sorbent that must be determined from isotherm models. For this work, Freundlich
and Langmuir models have been chosen corresponding to the forms shown in Eqs.
(2.4 and 2.5) to represent the experimental measurements of sorption process after
reaching the equilibrium state. Defining the appropriate isotherm model based on
the familiar statistical measures like R2, SSE, RMSE and X2 was essential for
calculating the retardation factor in the advection-dispersion equation which must
be applied to describe the contaminant transport in the subsurface environment as
in eqs. (2.7-2.9). The high values of R2 and low values of nonlinear statistical
parameters such as SSE, RMSE and X2 which indicate a good fitting model to
experimental data for each antibiotic (Marzbali & Esmaieli, 2017; Wang & Guo,
2020).

The constants of applied isotherm models (Table 4.4) were calculated using
nonlinear regression through application of “solver” option within Excel software
2016 (Faisal and Naji , 2019). Based on the highest values of R 2 and small values
of SSE, Langmuir model has high ability in the representation the sorption of TC
and AMX on the prepared sodium alginate beads. The matching between the
experimental measurements and sorption isotherms can be observed from Figs.
(4.20) and (4.21); however, this concurrence also can recognize from highest R 2 (≥
0.9834) and (≥ 0.9486) and lowest values of SSE, RMSE and X 2 parameters for
Langmuir model were less than Freundlich model as (1.358, 0.3514 , 43.527 ) for
TC and (5.067 , 0.6787 , 0.5424) for AMX antibiotic.

Whether the adsorption is favorable at the examined conditions can be


evaluated based on the two a aforementioned criteria. Separation factor (R L) and
adoption intensity parameter (n) derived from the Langmuir and Freundlich
models, respectively. It was noticed that all the RL values for TC and AMX within
the range (0–1) and this indicate Langmuir isotherm is favorable for sorption TC
and AMX onto (Ca/Fe)-LDH-sodium alginate beads. This is the standard case
when adsorption occurs normally, which is not so strong, but noticeably occurs,
and we can observe the shape of the adsorption isotherm. In the case of the
completely ideal irreversible case, the degrees tended toward zero rather than unity
which correspond to a completely reversible case (Li et al., 2022). The maximum
capacity of TC and AMX sorption has value of 10.393 and 13.322 mg/g
respectively, which means that the alginate beads have high ability in the
elimination of TC. Which indicated that the molecular adsorption for both
antibiotics occurred as monolayer and homogeneous (Togue, 2019).

Table (4.4). Outputs of fitting process for isotherm models in the sorption of
TC onto sodium alginate beads.
Figure (4.20): Models of sorption isotherms for interaction of prepared
sorbent - water contaminated with TC.

Figure (4.21): Models of sorption isotherms for interaction of designed sorbent -


AMX water contaminated

To compare the present method with other methods, adsorption capacity,


pH, contact time and initial TC and AMX concentrations are all depended in this
comparison depending on above results and this is provided in Tables (4.5) and
(4.6). The present method of synthesis (Ca/Fe)-LDH-sodium alginate obtained
accepted adsorption capacity with shorter equilibrium time indicating the better
performance of the prepared nanomaterial from egg shell waste for TC and AMX
antibiotics removal due to adsorption mechanisms. Freundlich constant (K f) was
3.0711 and 6.557 (mg/g) (L/mg) 1/n
for TC and AMX onto synthesized adsorbent
beads while (n) > 1 for both antibiotics ; thus, the curves of isotherm can classify
as “favorable” ( Putra et al. ; Al-Ghouti & Da'ana ,2020).

Table (4.5). Comparison the TC adsorption capacity of this work with


previous studies.

Table (4.6). Comparison the AMX adsorption capacity of this work with
previous studies.

4.8 Predominant mechanisms


The adsorption mechanism analysis is essential because it can give a
theoretical basis for the usage of (Ca/Fe)-LDH-sodium alginate in certain antibiotic
removal methods, as well as highlight its effectiveness as applied it in field area.
The general mechanisms for antibiotic adsorption onto adsorbents include
electrostatic interaction (cation and anion attractions), hydrogen bonding, surface
precipitation, intra-particle diffusion, and π-π interactions (Tan et al., 2015). Four
processes are involved in the adsorption of contaminants firstly: solute
transportation in bulk, secondly: adsorbate film diffusion, then adsorbate pores
diffusion, and finally: adsorption, which is the interaction of the adsorbate with the
porous structure. These interactions, which are restricted to the monolayer
covering, are stronger and more focused than physical adsorption (Ahmed et al.,
2015).

As a result, mechanisms were explored based on sorbent surface chemistry


and structure using FT-IR spectrum, SEM-EDS, and XRD analysis before and after
antibiotics adsorption and it’s illustrated in figure (4.22). Because cavities with
protruding and receding surfaces increased antibiotic adsorption (Fan et al., 2022)
with various adsorption mechanisms, the characteristics of adsorptive materials
have a significant influence in the adsorption of organic molecules. and a
significant differences in the morphology of adsorbent beads after the sorption of
TC and AMX antibiotics molecules onto sorbent were seen when compared to the
morphology of the sorbent surface before to the sorption process due to
aggregation and filling with antibiotics molecules (Mansour et. al., 2022).

The survey EDS revealed that the element percentages C, Na, N, O, S, Fe,
Ca, and H in the composition of (Ca/Fe)-LDH-sodium alginate beads can change
and increase in magnitude after the sorption process, as shown in figure (4.7). The
remarkable increase in C and N content explains that occurs after TC sorption,
which confirms the adsorption of TC antibiotic. The same trend was seen in the
EDS analysis of adsorbent beads after loading with AMX; the increase in C, N,
and S elements indicates that antibiotic molecules and prepared (Ca/Fe)-LDH-
sodium alginate beads interact successfully. The overlapping of (-OH) and amides
(N-H) stretching vibration bands produced the typically broad and strong band
extending from (3600 to 3100) cm-1 after adsorption. While the asymmetrical and
symmetrical stretching vibrations of the (-C-H) bonds did not appear because they
overlapped with the strong and broad stretching vibrations of the hydroxyl and
amides (N-H) groups and the absorption peak at 3317 cm -1 for both TC and AMX.
This results from the electrostatic interaction between TC and AMX molecules and
hydroxide layers to achieve their safe stabilization in the interlayer space of LDH
(Zhang et sl., 2019).

The appearance of characteristic S=0 peak at frequencies (1415 cm -1 to


1380 cm-1 ) and (1070 cm-1 to 1030 cm-1) due to the sulfate and sulfoxide groups at
AMX-loaded adsorbent beads ,can provided a significant evidence of amoxicillin
molecular adsorption onto the (Ca/Fe)-LDH-sodium alginate adsorbent with
supporting of mentioned groups.(Ragab et. al., 2021). These variations indicated
that the (-OH), (-COO-), (-C-O), and (-C-C-) bonds could play a role in the
antibiotic adsorption process via hydrogen bonding (Kang et al., 2010; Zhuang et
al., 2017).

Because of the presence of benzene rings in the antibiotics, these alterations


indicated that these active functional groups participated in the adsorption of
tetracycline and amoxicillin on the adsorbent beads by corresponding active
oxygen-containing functional groups via the strong (n-π EDA interaction) and
(hydrogen bonding) (Chen et al., 2013). Adsorption mechanism will be better
understood by investigating the properties of both adsorbent beads and antibiotics
that effected by different parameters such as solution pH, temperature (Pan and
Xing, 2008).

Sorption of antibiotics on beads at low pH (3-6) was controlled by the( π-π


EDA interaction) as driving force with adsorbent surface rather than electrostatic
force, therefore, the removal ratios for TC and AMX were low at acidic pH. It's
also possible that the competition between the free H + ions in solution and the
protonated antibiotic functional groups (carboxyl and amine groups) was the main
factor causing this behavior, also the interactions between protonated groups in
(Ca/Fe)-LDH-sodium alginate beads surface and negatively charged oxygen atoms
of carboxyl groups when the solution pH lies below its pHpzc due to the presence
of free H+ ion in the solution (Putra et al., 2009).

In other hand, the protonated functional groups found in antibiotics that have
a positive charge exchanged low valence cations (Fe+3 and Ca+2) in an adsorbent
structure, and this is made clear by the decrease in the percentage of these cations
in the solution after adsorption is completed. when pH equal to pHpzc at value =7 ,
adsorbent beads contain both positively and negatively charged functional groups
at the same time which attract both positively and negatively charged functional
groups on antibiotics, this led to the maximum removal performance from
adsorbent beads which enhanced by the electrostatic force , and this equal a
decrease of zeta potential from range (2-6) so there was increase in the adsorption
of antibiotics onto Ca/Fe-LDH-Na-alginate beads and that is agreed well with the
results above (Dutta et al., 1999 ; Álvarez-Esmorís et al., 2022).

The hydration and ionization of antibiotic can decrease at neutral condition


and this can enhance the removal process through the hydrogen bonds and π– π
stacking effect. When the pH further increased, the adsorption capacity was
decreased, where it refer to pH > pH pzc and the adsorbent surface become
negatively charged and combined with H+ groups of antibiotics leading to the
electrostatic repulsion between the charge of beads surface and OH- groups of
antibiotics. Also, the competition between antibiotics and surface functional groups
e.g. –COOH and –OH can be another reason for decreasing in the sorption at
higher pH, thus other experiments for TC and AMX adsorption by this material
were conducted at pH 7. The surface complexes between the negatively charged
groups of antibiotics and the positively charged hydroxyl groups (-Fe-OH or -Ca-
OH) on beads, which would be one of the main adsorption mechanisms as
indicated by (Harja and Cioban, 2018).

With no an additional absorption bands that observed in the FTIR analysis


after antibiotic adsorption; therefore, the binding does not involve the formation of
new covalent bonds between antibiotic molecules and the functional groups of
adsorbent beads and indicating only a superficial complexation mechanism. By
studying the kinetic, thermodynamic and isotherm data, it was assumed that the
adsorption occurs for both antibiotics normally as monolayer and homogeneous
onto (Ca/Fe)-LDH-sodium alginate beads. Then the reaction between antibiotics
molecules and adsorbent surface was a chemisorption same as most of adsorbed
antibiotics with two binding sites according to the pseudo second order kinetic
model (Ahmed et al., 2015).as shown as in figure .

Figure (4.22): Adsorption mechanisms for TC and AMX antibiotics


on (Ca/Fe)-LDH-sodium alginate bead

4.9 Reusability and Regeneration Study


Reusability represents an important aspect to evaluate the performance of
adsorbents and to evaluate its potential for practical application because the
reusability can obviously reduce the operation cost and make the process
economical. With the adsorption-desorption test reliability of (Ca/Fe)-LDH-
sodium alginate beads has been investigated in this paper to achieve the required
regeneration for this sorbent and conducted to evaluate the possibility of reuse the
exhausted (Ca/Fe)-LDH-sodium alginate beads for the removal of TC and AMX.
In order to establish the reusability of adsorbent bead, sequential cycles were
repeated five times with the same batch conditions. The percentage removal of TC
and AMX were found to be 95.2% and 91.4 % in first adsorption cycle whereas the
reused beads shows capacity around 93.35% and 88.27% respectively at the end of
2nd a cycle. Fig.(4.23) presents the removal percentages of both antibiotics as a
function for the number of reuse times.

Figure (4.23): Reusability of (Ca/Fe)-LDH-sodium alginate for sorption of TC and


AMX antibiotics

This figure signifies that there is decreasing in the removal efficiencies with
an increase of reuse times number and the achieved removal efficiencies after 5
cycles exceeded 73.53% and 68.98 %; and the observed loss of sorption
performance during the long-term use might be due to the blockage of some active
sites and by changes in the chemistry and the structure of the sorbents
(Vijayalakshmi et al., 2017).

In order to examine the recyclability of the adsorbent, three regenerating


methods were examined. First ,the adsorbent was regenerated using 10 ml of 0.1 M
sodium hydroxide (NaOH) to desorb TC and AMX from and regenerated the
exhausted sorbent that originally applied in the batch test of sorption at the best
operational conditions with agitation speed 200 rpm for 90 min at room
temperature. The beads were washed with deionized water subsequently. After
drying, the next adsorption cycle was carried out. A total of four cycles of
regeneration were taken, and the TC and AMX removal percent after each cycle
was calculated. This procedure also repeated for regeneration adsorbent but by
using 10 ml of 0.1 M hydrochloric acid (HCl) and distilled /deionized water only
for four runs.

Fig. (4.24) shows that when the (Ca/Fe)-LDH-sodium alginate beads was
regenerated using sodium hydroxide, the adsorption removal efficiency for TC and
AMX could reach (80.96 %) and (77.822%) after four runs , comparable with that
from first run which is (95.89 %)and ( 93.34 %) . While when using hydrochloric
acid (HCl) and distilled /deionized water the TC removal efficiencies reach
(77.43% ) and (83.131%) respectively after four runs against from first run values
(94.83 %) and (95.851 %) , while the AMX removal efficiencies were at the end
of recycling with HCl (73 % ) and distilled /deionized water (75.33 %) from what
it equal at initial cycle values (92.433 %) ,(93.12 %) and (%)with HCl and D.water
respectively.

Figure (4.24): Removal efficiency changes for a).TC and b).AMX onto (Ca/Fe)-
LDH-sodium alginate beads after four regeneration cycle

The removal efficiency for TC did not reduce below 77 % or 73% for AMX
after a 4th. regeneration cycle, suggesting the adsorbent was reusable under either
regenerating conditions with minimum structural change occurring as shown in
figure (4.25).The findings indicated that (Ca/Fe)-LDH-sodium alginate had good
reusability, and it can be adopted for TC and AMX adsorption in the practical
conditions and these efficiencies improving by regeneration process as shown in
figures.

Figure (4.25): (Ca/Fe)-LDH-sodium alginate beads deformation after four


regeneration cycle

4.10 Adsorption Performance in Fixed Bed Column Experiments

Basic kinetic data in the form of breakthrough curves are essential for
developing and optimizing adsorption operations. The test was carried out on a
column packed with (Ca/Fe)-LDH-sodium alginate beads, with the duration
extended for 380 and 360 hours for TC and AMX, respectively. The breakthrough
curve of an adsorbate in a continuous system is expressed by charting the outlet to
inlet concentration (C/Co) ratio versus time (t) or throughput volume and this curve
describes the dynamics of a continuous adsorption system, whose behavior is
connected to the shape of the adsorption isotherm and is influenced by the
diffusional stages inside the fixed bed. The gathered data were then applied
utilizing mathematical equations of the column system. Breakthrough occurs at
contact time where C/Co = 0.05 and a typical breakthrough curve includes the mass
transfer zone (MTZ) where adsorption takes place. The following figures (4.26-
4.45) and tables (4.7-4.16) demonstrate the breakthrough curves and operation
conditions for antibiotics adsorption onto (Ca/Fe)-LDH-sodium alginate beads
respectively for the selected mathematical models for the fixed-bed systems. The
breakthrough profiles had the expected smooth S shape, which is governed by the
shape and length of the mass transfer zone MTZ and the continuous antibiotics
biosorption process was continued until the column reached saturation, i.e. until
the output concentration was at least 0.9 Co.

The results of the breakthrough curves was obtained with for the flow rates
(Q) of 5, 10 and 15 mL/min, bed heights (L) of 5,10,15,20, and 25 cm, and initial
concentrations (Co) of 20, 50 and 100 mg/L, to evaluate the reactivity of (Ca/Fe)-
LDH-sodium alginate beads in the remediation process through monitoring the
antibiotics concentrations in the effluent from port P1.2.3.4. and 5. From tables of
parameters for TC antibiotic breakthrough curves, it’s obtained that the Bohart-
Adams model (R2 = 0.993800-0.999524), Thomas model (R2 = 0.993800-
0.999134), Bulter-Cussler-Hu model (R2 = 0.931981-0.974377) and Clark model
(R2 = 0.862154-0.99939) , and for AMX antibiotic it was (R2 = 0.993156-
0.999208) for the Bohart-Adams model, (R2 = 0.993021-0.999208) for Thomas
model, while Bulter-Cussler-Hu model (R2 = 0.95464-0.983606) and Clark model
(R2 = 0.995205-0.999166) , exhibit higher determination coefficients than the Yan
model as (R2= 0.83559-0.8991099) and (R2= 0.830211-0.966229) for TC and
AMX respectively.

Unlike the Bohart-Adams, Thomas, Bulter-Cussler-Hu and Clark models,


the Yan model might provide an appropriate description only for the initial part of
the breakthrough curve; therefore, it was less convenient when used to describe the
breakthrough behavior compared to the other models.The purpose of this process is
to predict the performance of the adsorbent and its capacity for removing
contaminants from aqueous solutions before applying it in the field because the
column study is an ease operation, faster process, and simplicity of scale-up
process

4.10.1. Effect of Contaminants Initial Concentration

The function of an adsorption column was investigated for various TC and


AMX antibiotic input concentrations (20, 50, and 100 mg/L) at flow rates equal 5
mL/min for the ports P1 to P6 with bed heights of 5, 10, 15, 20, and 25cm. The
effect of initial TC and AMX concentrations on the process of antibiotic adsorption
was represented in the breakthrough curves shown in (Figures 4.26–4.30) and
(Figures 4.31–4.35), respectively.

Tables (4.7-4.12) show that changing the concentration gradient had a


significant effect on the breakthrough and exhaustion times, and that the
equilibrium uptake increased with increasing initial concentrations, whereas the
breakthrough and exhaustion times remarkably decreased with increasing inlet
antibiotic concentrations, because the adsorbing sites on the adsorbent surface
became quickly saturated because adsorbent particles were exposed to higher
levels of adsorbed materials ; therefore, as concentration increased, sharper
breakthrough curves were detected, indicating a smaller mass transfer zone and a
faster adsorption rate (Rafati, et al., 2019).

For example ,the breakthrough time for bed high of 25 cm (P5) at the
influent TC concentration of 20 mg/L and flow rate=5 mL/min appeared after 102
hr. while the breakpoint time to the same port decreased and occurred after 78 and
49 hr. at the influent TC concentration of 50 and 100 mg/L respectively. As a
result, it is easy to observe from the figures that the column performance was fast
at first and remained constant at equilibrium after the removal efficiency began to
decrease, while at lower influent concentrations, the breakthrough occurred
gradually and the breakthrough curves scattered and changed to a sharper curve as
the influent concentration increased. Generally, a higher influent concentration
results in a higher contaminants adsorption, because they would have a greater
driving force on the adsorbent surface. This could occur because a higher
concentration gradient causes a greater driving force for the mass transfer of
antibiotics molecules because the large difference between the antibiotics in the
solution and the antibiotics on the surface of the adsorbent, and thus the adsorption
sites are covered more quickly (Vu, et al., 2018).

The lower concentration resulted in greater column performance since the


decrease in TC and AMX concentrations meant that a larger volume of the solution
could be treated before exhaustion. In contrast, the Thomas adsorption capacity for
TC and AMX increased from (133.92 to 264.6) mg/g and from (148.8 to 354)
mg/g, when the influent concentration increased in port 5 from 20 mg/L to 50
mg/L, respectively at Q=5 mL/min. Also, it was observed that the adsorption
capacity was not proportionally increased with increasing influent concentration.
As a result, adsorption capacities of antibiotics were higher when inlet
concentration increased This can be explained by the fact that the greater
concentration gradient led to a faster mass transfer due to the increased surface
diffusion coefficient (Marzbali & Esmaieli , 2017).

Figure (4.26): Breakthrough curves models in comparison with experimental data


for different initial concentrations of TC antibiotic for Port1 at Q=5 mL/min.

Figure (4.27): Breakthrough curves models in comparison with experimental data for
different initial concentrations of TC antibiotic for Port 2 at Q=5 mL/min.

Figure (4.28): Breakthrough curves models in comparison with experimental data for
different initial concentrations of TC antibiotic for Port 3 at Q=5 mL/min.

Figure (4.29): Breakthrough curves models in comparison with experimental data for
different initial concentrations of TC antibiotic for Port 4 at Q=5 mL/min.

Figure (4.30): Breakthrough curves models in comparison with experimental data for
different initial concentrations of TC antibiotic for Port 5 at Q=5 mL/min.
Table (4.7): Parameters of mathematical models from fitting the TC breakthrough curves
data at ports 1,2,3,4 and 5 for 20 mg/L inlet concentration and Q=5mL/min

Table (4.8); Parameters of mathematical models from fitting the TC breakthrough curves
data at ports 1,2,3,4 and 5 for 50 mg/L inlet concentration and Q=5mL/min

Table (4.9); Parameters of mathematical models from fitting the TC breakthrough curves
data at ports 1,2,3,4 and 5 for 100 mg/L inlet concentration and Q=5mL/min

Figure (4.31): Breakthrough curves models in comparison with experimental data for
different initial concentrations of AMX antibiotic for Port 1 at 5 mL/min.

Figure (4.32): Breakthrough curves models in comparison with experimental data for
different initial concentrations of AMX antibiotic for Port 2 at 5 mL/min.

Figure (4.33): Breakthrough curves models in comparison with experimental data for
different initial concentrations of AMX antibiotic for Port 3 at 5 mL/min.

Figure (4.34): Breakthrough curves models in comparison with experimental data for
different initial concentrations of AMX antibiotic for Port 4 at 5 mL/min.

Figure (4.35): Breakthrough curves models in comparison with experimental data for
different initial concentrations of AMX antibiotic for Port 5 at 5 mL/min.
Table (4.10): Parameters of mathematical models from fitting the AMX breakthrough
curves data at ports 1,2,3,4 and 5 for 20 mg/L inlet concentration and Q=5mL/min

Table (4.11): Parameters of mathematical models from fitting the AMX breakthrough
curves data at ports 1,2,3,4 and 5 for 50 mg/L inlet concentration and Q=5mL/min

Table (4.12): Parameters of mathematical models from fitting the AMX breakthrough
curves data at ports 1,2,3,4 and 5 for 100 mg/L inlet concentration and Q=5mL/min
4.10.2. Effect of the solution flow rate

The effects of the solution flow rates (Q; mL/min) on the breakthrough
curves for antibiotics adsorption were studied at a bed height of 5,10,15,20 and 25
cm and initial TC and AMX concentrations of 20 mg/L as in figures(4.36- 4.45).
Maintaining a low flow rate in the adsorption system increases the contact time
between antibiotics molecules and (Ca/Fe)-LDH-SA beads and consequently
increases the removal of antibiotics. The breakthrough and exhaustion times for
TC at port 5 in fig. (4.40) significantly decreased (from 102 to 84) hr. and from
(252 to 228) hr., respectively, as the flow rate increased from 5 mL/min to
10mL/min, respectively. While for AMX in fig. (4.45) it’s decreased at port 5
(from 88 to 83) hr. and from (272 to 260) hr., respectively at the same conditions.

As expected, the breakthrough curves of both antibiotics were steeper and


the breakthrough time decreased as the flow rate increased, because increasing
flow result in insufficient retention time of the antibiotics in the column and the
molecules have a minimum of contact time with adsorbent, thus reducing the
diffusion of the antibiotics molecule into the pore of the (Ca/Fe)-LDH-SA bead
(Rafati et al., 2019). Furthermore, the high flow rate may led to some of the
attached molecules of the solute to be desorbed through reversible and weak
bindings with the sorbent surface ; therefore, the contaminant concentration rapidly
increases, resulting in an early time of breakthrough (Thirunavukkarasu et al. ,
2021).

Because the antibiotic molecules had enough time to settle in the fixed-bed
column and had more opportunities to come into contact with the adsorbent, the
highest amount of adsorbed material was observed at a flow rate of 1 mL/min,
which led to a significant removal of antibiotic molecules from the column.
Accordingly, the relationship between flow rate and removal efficiency is inverse
(Vu et al., 2018).

Figure (4.36): Breakthrough curves models in comparison with experimental data for
different water flow rates of TC antibiotic at Ports 1 for Co=20 mg/L

Figure (4.37): Breakthrough curves models in comparison with experimental data for
different water flow rates of TC antibiotic at Ports 2 for Co=20 mg/L

Figure (4.38): Breakthrough curves models in comparison with experimental data for
different water flow rates of TC antibiotic at Ports 3 for Co=20 mg/L

Figure (4.39): Breakthrough curves models in comparison with experimental data for
different water flow rates of TC antibiotic at Ports 4 for Co=20 mg/L

Figure (4.40): Breakthrough curves models in comparison with experimental data for
different water flow rates of TC antibiotic at Ports 5 for Co=20 mg/L

Table (4.13): Parameters of mathematical models from fitting the TC breakthrough curves
data at ports 1,2,3,4 and 5 for flow rate Q=10 mL/min and 20 mg/L inlet concentration

Table (4.14): Parameters of mathematical models from fitting the TC breakthrough curves
data at ports 1,2,3,4 and 5 for flow rate Q=15 mL/min and 20 mg/L inlet concentration
Figure (4.41): Breakthrough curves models in comparison with experimental data for
different water flow rates of AMX antibiotic at Ports 1 for Co=20 mg/L

Figure (4.42): Breakthrough curves models in comparison with experimental data for
different water flow rates of AMX antibiotic at Ports 2 for Co=20 mg/L

Figure (4.43): Breakthrough curves models in comparison with experimental data for
different water flow rates of AMX antibiotic at Ports 3 for Co=20 mg/L

Figure (4.44): Breakthrough curves models in comparison with experimental data for
different water flow rates of AMX antibiotic at Ports 4 for Co=20 mg/L

Figure (4.45): Breakthrough curves models in comparison with experimental data for
different water flow rates of AMX antibiotic at Ports 5 for Co=20 mg/L

Table (4.15): Parameters of mathematical models from fitting the AMX breakthrough
curves data at at ports 1,2,3,4 and 5 for flow rate Q=10 mL/min and 20 mg/L inlet
concentration

Table (4.16): Parameters of mathematical models from fitting the AMX breakthrough
curves data at ports 1,2,3,4 and 5 for flow rate Q=15 mL/min and 20 mg/L inlet
concentration

4.10.3. Effect of Bed Height (Used Bio-adsorbent Mass)

The study on the influences of (Ca/Fe)-LDH-SA beads bed height on the process
of ammonium adsorption was performed at initial TC concentration of 20 mg/L
and flow rate of 5 mL/min. Breakthrough curves of antibiotics adsorption for
example on the adsorbent at different bed heights (5, 10, 15, 20 and 25 cm) and
The results showed that the breakthrough and exhaustion times for TC decreased
as follows: 37 hr. and 124 hr. at Z = 5 cm, 55 hr. and 160 hr. at 10 cm, 80 hr. and
190 hr. at 15 cm, 95 hr. and 212 hr. at 20 cm and 101 hr. and 252 hr. at 25 cm,
respectively. We can notice that the breakthrough time for TC can increase
dramatically due to increase of bed depth from 5 cm (P1) to 25 cm (P5)
respectively, while this time is also increased from 51 hr. and 144 hr. at Z = 5 cm,
to 58 hr. and 170 hr. at 10 cm, while it equal to 66 hr. and 210 hr.at 15 cm, 74 hr.
and 214 hr. at 20 cm and 86 hr. and 272 hr.at 25 cm, respectively under the same
conditions for AMX antibiotic.

Because the breakthrough curve slope becomes smaller as bed height


increases, a wider mass transfer zone results. This increases the interaction
between antibiotics and the active sites on (Ca/Fe)-LDH-SA beads by increasing
the breakthrough and exhaustion times at the higher port (Altufaily et al., 2019).
Furthermore, a higher used (Ca/Fe)-LDH-SA beads mass (bed height), augmented
the total surface area of adsorbent and consequently more available active sites for
the antibiotics adsorption. As a result, the highest adsorption capacity for TC and
AMX (133.92 and .148.8) mg/g were obtained at a bed height of 25 cm and the
lower values were (73.2 and 82.8) mg/g at 5 cm height, where the increase in
adsorbent mass could improve the column operation and reduce the effluent
antibiotics concentration at the end of the system (Vu et al., 2018).

Additionally, due to both breakthrough and exhaustion times were increased


with increasing the bed height, this contributed the beds were saturated in a longer
time by a larger adsorbent mass (Liao et al., 2013; Li et al., 2018).

You might also like