Next Article in Journal
A Laboratory Set-Up for the Analysis of Intermittent Water Supply: First Results
Previous Article in Journal
Distribution and Transfer of Antibiotic Resistance Genes in Coastal Aquatic Ecosystems of Bohai Bay
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Granular Natural Zeolites: Cost-Effective Adsorbents for the Removal of Ammonium from Drinking Water

Institute for Water Chemistry, Technische Universität Dresden, 01062 Dresden, Germany
*
Author to whom correspondence should be addressed.
Water 2022, 14(6), 939; https://doi.org/10.3390/w14060939
Submission received: 17 February 2022 / Revised: 10 March 2022 / Accepted: 14 March 2022 / Published: 17 March 2022
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
Increasing food demand has resulted in an ever increasing demand for nitrogen fertilizers. Ammonium is the main constituent of these fertilizers and is a threat to aquatic environments around the world. With a focus on the treatment of drinking water, the scope of this study was to investigate the influence of key parameters on the suitability of granular natural zeolites as adsorbents for ammonium. Sorption experiments were performed in artificial matrices by varying the grain size, contact time, ammonium concentration, pH, content of competing ions, and regeneration solutions used. Additionally, natural matrices and the point of zero charge (pzc) were investigated. With an initial ammonium concentration of 10 mgN/L, the grain size was shown to have no significant effect on the sorption efficiency (97–98%). The experimental data obtained was best described by the Langmuir adsorption model (R2 = 0.99). Minor effects on sorption were observed at different pH values and in the presence of competing anions. In addition, the pHPZC was determined to be between pH 6.24 and pH 6.47. Potassium ions were shown to be better than sodium ions for the regeneration of previously loaded zeolites, potassium is also the main competitor to ammonium sorption. The use of tap, bank filtrate, river, and groundwater matrices decreased the ammonium sorption capacity of granular natural zeolites by up to 8%. Based on our results, granular natural zeolites are promising cost-effective adsorbents for drinking water treatment, especially in threshold and developing countries.

1. Introduction

The increase in industrialization that has occurred since the early 18th century has resulted in a 16-fold increase in the living standard of the average world citizen [1]. Concomitantly, the world’s population has risen from 1 billion to 7.7 billion between 1800 and 2019 [2,3]. Inevitably, an increase in food production was necessary to support this population increase. To enable food production on such a tremendous scale, the use of fertilizers in agro-cultural applications is essential. The development of an energy-intensive ammonia synthesis method (10 kWh/kg NH3) [4] by Fritz Haber and Carl Bosch in the early 20th-century led to nitrogen fertilizer production increasing by 776% between 1961 and 2014 [5].
The excessive use of nitrogen fertilizers has enormously accelerated the global nitrogen (N) cycle and has resulted in an omnipresent form of soil and water pollution [6,7,8,9]. Ammonium (NH4+) is the main constituent of nitrogen fertilizers and, for this reason, is one of the principle concerns to address; other concerns include ammonification, septic systems, landfills, animal wastes, irrigation, or municipal and industrial wastewater [10]. Further compounding the issue, high N concentrations in surface and groundwater lead to additional concerns: increased oxygen demands, stimulation of eutrophication, soil acidification, and others. Broadly speaking, the examples of NH4+ removal discussed in the literature can be divided into two types: removal from wastewater and removal from drinking water. These examples differ fundamentally in terms of their characteristics, particularly their pollutant loads.
The pollutant loads, including ammonium levels, tends to be much higher in wastewater applications: (1) generally 30 mgN/L in municipal wastewater [11,12]; (2) up to 1500 mgN/L in anaerobic digestion liquor from sludge treatment [13]; and (3) up to 6000 mgN/L in high strength ammonium industrial wastewater [14].
On the other hand, drinking water sources are often derived from groundwater and bank filtrate. They tend to exhibit lower pollutant loads due to various elimination processes (e.g., microbial conversion, adsorption) during soil passage. Usually, the natural NH4+ concentration in aerobic groundwater and surface water is <0.2 mg/L. However, under anaerobic groundwater conditions, the concentration may increase more than tenfold [15]. For instance, a number of sample sites have been documented to contain concerning levels of ammonium: 14 mg/L in the shallow groundwater of the Jianghan Plain, China [16]; up to 28 mg/L in the bank filtrate of Reijerwaard along the river Nieuwe Maas, Netherlands [17]; up to 100 mg/L in groundwater systems in the metropolitan area in Hanoi, Vietnam [18,19]; as well as up to 390 mg/L in a coastal aquifer-aquitard system in the Pearl River Delta, China [20]. Recent investigations have also reported a significant correlation between high NH4+ concentrations in groundwater and high contents of dissolved iron and manganese [21]. Considering that the world population’s food demands continue to increase, it is reasonable to assume that NH4+ levels will also increase in various drinking water matrices.
In terms of human toxicity, NH4+ only becomes harmful over 100 mg/kg body weight per day [15]. However, during drinking water treatment processes, the formation of toxic nitrite (NO2) and harmful chlorination side products, as well as a higher chlorine consumption, must be considered. Generally, NH4+ represents a pollution indicator for drinking water and must not exceed specific threshold values following purification (e.g., 0.5 mg/L in Germany) [22]. Therefore, NH4+ contaminated waters can be treated by various technologies and processes including: (1) activated sludge and fixed bed reactor processes; (2) air-stripping; (3) membrane processes; and (4) ion-exchange by zeolites [23,24].
When considering the treatment processes highlighted above, natural zeolites, particularly clinoptilolite, stand out as promising adsorbents for drinking water treatment. The use of zeolites is more economical because high energy expenditures are avoided by excluding aeration processes and because additional chemicals and high pressures are not required. Furthermore, the use of zeolites is sustainable; ammonia synthesis can be reduced since NH4+-saturated particles can be repurposed as slow-release fertilizers in agriculture [25] or can be regenerated in a cost-effective manner using sodium solutions [26,27,28]. Another positive feature of natural zeolites is their abundant availability and their low cost. Approximately 1 million tons of natural zeolites were mined globally in 2020, with the major producers being China, South Korea, Indonesia, and Slovakia [29]. Mining costs are generally low and achieving the desired zeolite reduction ratio during the milling process is the main expense. For example, the price of untreated finer zeolite products is USD 50 to 120 per ton and twice as high as for granular products, at USD 30 to 70 per ton [30,31]. In summary, outstanding NH4+ removal efficiencies, the convenience of application, their sustainable characteristics, and their broad range of applications means that natural zeolites hold tremendous promise as highly cost-effective adsorbents [32,33,34].
Zeolites are constructed by a three-dimensional tetrahedral structure of AlO4 and SiO4, covalently connected by an oxygen atom to form interconnected channels within the zeolites’ framework. The higher the AlO4 content, the higher the negative surface charge. This surface charge is compensated for by adsorbed cations such as Na+, K+, Ca2+, and Mg2+, which can be exchanged for NH4+ with remarkable selectively [35].
Previous studies have investigated the removal of heavy metals (Zn, Cr, Pb, Cd, Cu, etc.) [36,37,38], organic contaminants [39], and high NH4+ loads by natural zeolites during wastewater treatment [39]. Natural zeolites have also been applied to the removal of NH4+, humic acid, Fe, and Mn from surface and groundwater matrices. Two similarities were identified by comparing both areas of study: (1) similar zeolite grain sizes (powder-like and fine particles); and (2) zeolites were investigated in artificial, wastewater, or surface and groundwater matrices. In general, the composition, modification and quantity of the used zeolite, the initial concentration of the contaminants, pH, temperature, and contact time were identified to have the strongest influence on the removal efficiencies for various contaminants.
Until now, there has been little research on the effect of different water matrices on the efficiency of NH4+ removal using the same natural zeolite. Furthermore, to the best of our knowledge, granular natural zeolites have not been investigated for NH4+ removal. Finally, a trend can be seen when combining sorption onto granular natural zeolites and biological treatment in flow-through fixed bed reactors to enhance NH4+ removal [40,41,42]. From a process engineering viewpoint, the clogging of filters in flow-through fixed-bed columns is less likely to occur when using granular natural zeolites, thus improving filter runtime, and increasing intervals between backwashing.
The main objective of this study was to demonstrate the suitability of cost-saving granular natural zeolites (clinoptilolite) in terms of their NH4+ removal efficiency and to compare them directly with the finer grain sizes usually used in drinking water treatment. In a departure from previous studies, multiple water matrices such as artificial, as well as surface and river water, groundwater, and bank filtrate were investigated. Various process parameters were investigated for future applications of larger zeolite grain sizes, by conducting a broad range of batch experiments. This investigation demonstrated that the use of granular natural zeolites is more cost-effective than the use of finer zeolites. By avoiding milling costs and combining this with high NH4+ removal efficiencies and sustainable characteristics (e.g., N re-use for agriculture), granular natural zeolites are a highly promising route to improving the quality of drinking water, especially in threshold and developing countries.

2. Materials and Methods

2.1. Zeolite Characteristics

The natural zeolite (clinoptilolite) investigated, CLP85+, was supplied by Zeolith Umwelttechnik Berlin GmbH, Berlin, Germany. In Table 1, the chemical composition and general characteristics of the zeolite are shown. In order to remove particulate matter, salts, and excess adsorbed cations, the zeolites were washed with ultrapure water until the electrical conductivity was less than 10 µS/cm (water appeared clear by visual inspection). The zeolites were subsequently dried at 80 °C for 24 h before beginning the experiment.

2.2. Batch-Set-Up

Batch experiments to investigate the impact of various process parameters on the NH4+ sorption capacity of zeolites were carried out in triplicate at 22 °C. Matrices used included: ultrapure water; tap water; surface water; groundwater; and bank filtrate. The initial NH4+ concentration of 0.71 mM (12.8 mg/L; 10 mgN/L) for all experiments—except the investigations into ammonium load—was obtained using a 0.03 M NH4Cl (analytical grade) stock solution. For investigating the influence of contact time, three zeolite grain sizes (1–2.5 mm, 8–16 mm, 16–32 mm) were used. All other experiments were conducted exclusively using zeolite granules with a grain size of 8–16 mm. To avoid microbial contamination, the water matrices used, as well as the batch set-ups, were autoclaved prior to experiments. The batch set-ups were performed using the bottle point method and were carried out in borosilicate flasks (Borosilicate 3.3 glass; VWR International GmbH; Dresden, Germany) with 100 g of zeolites per liter of water [44]. Flasks were stirred at 50 rpm (SM-30 orbital shaker; Edmund Bühler GmbH; Bodelshausen, Germany), and sampling was performed at defined intervals until equilibrium was reached. The samples were filtered through a 0.45 µm PET filter (CHROMAFIL® Xtra PET-45/25; Macherey-Nagel GmbH and Co., KG; Düren, Germany), before determining the amount of residual NH4+.
To calculate the NH4+ sorption capacity (qt(NH4+)) and the normalized sorption capacity in equilibrium (normalized qeq) of zeolites, and to determine the sorption efficiency ESorption (%), the following equations were used:
q t ( NH 4 + ) =   C 0     C t m Z · V
normalized · q eq ( NH 4 + ) =   q Experiment q Control
E Sorption ( % )   = ( 1 q Experiment q Control ) · 100 %
The sorption capacity qt(NH4+) (mgNH4+/gZ) describes the time-dependent amount of adsorbed NH4+ ions per unit weight of zeolite (gZ). C0 and Ct are the initial and time-de-pendent NH4+ concentrations (mg/L) in solution. mZ (gZ) and V (L) are the adsorbent mass and the volume of treated water, respectively. qExperiment and qControl are the sorption capacities of a specific experiment and of an ultrapure water control, respectively. By using an ultrapure water control, a mostly non-competitive water matrix should be provided. qExperiment was normalized against qControl to get the normalized qeq(NH4+) (−). Before and after an experiment, the pH, electrical conductivity, and water temperature were recorded.

2.2.1. Influence of Contact Time and Ammonium Load

The influence of contact time was investigated under time-dependent sampling from 5 min to 20,160 min (14 days). To analyze the sorption kinetics in more detail, the experimental data were fitted against pseudo-first order reaction, pseudo-second order reaction, Elovich, and intra-particle diffusion kinetic models using the following equations [45]:
Pseudo - first   order   reaction :   ln ( q eq     q t ) q eq   = k 1 · t
Pseudo - sec ond   order   reaction :   t q t   = 1 k 2 q eq 2 + 1 q eq
Elovich :   q t = ( 1 b ) · ln ( a · b ) + 1 b · ln   ( t )
Intra - particle   diffusion :   q t = k i t 0 . 5 +   C
where qeq (mgNH4+/gZ) is the amount of adsorbed NH4+ ions per weight unit of zeolite at equilibrium: k1 (1/min) and k2 (g/mg·min) are the pseudo-first and second order rate constants and t is the contact time (min); a is the initial sorption rate (mg/g·min) and b describes the extent of surface coverage and activation energy (g/mg); ki is the intra-particle diffusion rate constant (mg/min0.5·g) and C is a constant related to the thickness of the boundary layer.
Additionally, the experimental data obtained was fitted to Langmuir and Freundlich adsorption models with an initial ammonium load of 1 to 5000 mgN/L. The Langmuir model (Equation (8)) supposes sorption at specific homogeneous sites within the adsorbent and assumes there is no interaction between adsorbate molecules. On the other hand, the Freundlich isotherm (Equation (9)) describes non-ideal sorption on heterogeneous surfaces and multi-layer sorption [44,46,47].
Langmuir   model :   q eq =   q max K L c eq 1 + K L c eq
Freundlich   model :   q eq = K F C eq 1 / n
qmax describes the maximum monomolecular sorption capacity (mg/g) and the constant KL (L/mg) the affinity between adsorptive and adsorbents. KF (L/g) characterizes the adsorption strength and n (−) determines the curve of the sorption isotherm.

2.2.2. Influence of pH

The influence of the solutions’ pH on the removal of NH4+ using a natural zeolite as an adsorbent was evaluated under the conditions specified in Table 2. A 0.1 M HCl solution was used to adjust the pH for pH 8 and 9. The buffer systems for pH 8 and 9 contained the greatest molar amount of sodium (due to the chemicals used in Table 2), which affects the ionic strength of the solution. To balance the background ionic strength, a 0.1 M NaCl solution was added equally to the other buffer systems. Finally, the 0.1 M buffer systems were diluted with ultrapure water to achieve a buffer concentration of 0.01 mM for the experiment. To ensure buffer stability, the pH was measured regularly.

2.2.3. Point of Zero Charge Determination

The point of zero charge (PZC) was determined by both the mass titration and pH drift methods. Differing from [48], a 0.03 M NaCl electrolyte solution as well as a reduced shaking speed of 50 U/min were used. By varying the mass concentration of zeolites between 0 and 112 g/L in the electrolyte solution, the change in pH was recorded over 24 h. The final pH values are plotted against the mass of zeolite per unit volume. The pHPZC is the pH that was achieved when the addition of further zeolite did not result in further changes in the pH. The pH drift method was performed by exposing the zeolites to the electrolyte solution and adjusting the pH between pH 2 to 9 by adding 0.1 M or 0.01 M HCl and NaOH solutions. After 24 h, the pH was measured again. The pHPZC is the pH value that no longer changes.

2.2.4. Influence of Competing Cations and Anions

Alkali metals and alkaline earth metals are ubiquitous cations in natural water matrices. To estimate the competing effect of Na+, K+, Mg2+ and Ca2+ ions on the NH4+ sorption capacity, concentrations of 0.7, 3.5, and 7 mM of each competing ion were investigated. To classify the possible effect of anions, the influence of NO2, NO3, SO42−, and HPO42−/H2PO4, each at a molar concentration of 3.5 mM, was investigated. Stock solutions with concentrations of 0.1 M NaCl, KCl, MgCl2·6H2O, CaCl2·6H2O, NaNO2, NaNO3, NaH2PO4, and Na2SO4 were diluted to obtain appropriate anion concentrations. It was also necessary to balance the background ionic strength using a 10 mM Na+ solution for the anion experiments (due to the chemicals used). All experiments with competing ions were normalized to a control that contained NH4+ in pure water.

2.2.5. Single and Multiple Regeneration

Zeolites have to be regenerated after a certain operating period. In this study, the zeolites were loaded with 10 mgN/L NH4+ in a first step and thereafter regenerated using 0.01 M and 0.1 M K+/Na+ solutions. Following the initial loading step, the zeolites were washed with ultrapure water until the electrical conductivity was less than 10 µS/cm and dried at 80 °C for 24 h to remove residual NH4+ from the surface and pores. After adding the regeneration solutions, time-dependent samples were taken and analyzed until equilibrium was reached. For multiple loading experiments, the zeolites were washed until the electrical conductivity was less than 10 µS/cm and dried at 80 °C for 24 h before another loading step was conducted. The cation concentrations before and after a loading step were also measured to identify any possible modifications of the zeolites’ chemical structure over time. The desorption efficiency can be calculated by the following Equation (10):
E Desorption ( % ) = ( 1 ( q Z · m Z c t · V Reg q Z m Z ) ) 100 %
where EDesorption is the desorption efficiency expressed as a percentage. qZ (mg NH4+/gZ) and mZ (g) describe the loading capacity of the previous loading step and the mass of zeolites used, respectively. The time-dependent NH4+ concentration (mg/L) and the volume of the regeneration solution (L) is described by ct and VReg.

2.2.6. Influence of Natural Matrices

For the investigation of real water matrices, Elbe river water, bank filtrate, groundwater, and a 1:10 mix of Elbe river water and tap water were analyzed. The NH4+ concentrations of the autoclaved water samples were measured and equal NH4+ levels were achieved by spiking with a 0.03 M NH4Cl stock solution. To estimate the effect of cations and the dissolved organic carbon (DOC) content on the loading capacity, the levels of both were measured before and after the experiment.

2.3. Analytical Methods

The standard parameters of pH, electrical conductivity, and temperature were recorded using a multimeter and sensors (Sentix®41 pH electrode, TetraCon®325 conductivity cell, Multi 340i multimeter; Xylem Analytics Germany Sales GmbH and Co., KG; Weilheilm, Germany). NH4+ was measured photometrically (Varian Cary® 50 UV-Vis Spectrophotomer; Agilent Technologies Corporation; Santa Clara, CA, USA) according to DIN 38 4606-E5-1. Two chromatography devices were used to measure the amount of cations (930 Compact IC; Methrom AG; Herisau, Switzerland) and anions (DionexTM ICS-6000; Fisher Scientific GmbH; Schwerte, Germany). DOC was determined by an analyzer (TOC-VCPN Analyzer; Shimazu Corporation, Kyoto, Japan).

3. Results and Discussion

3.1. Comparison of Different Zeolite Grain Sizes

3.1.1. Kinetic Evaluation

Initially, ammonium sorption onto zeolites was investigated at various time intervals for a 10 mgN/L NH4+ starting concentration. The influence of different zeolite grain sizes on the time-dependent sorption capacity of the zeolite qt(NH4+) are shown in Figure 1. Additionally, the initial rates (Figure 1a) and final stages (Figure 1b) are highlighted.
NH4+ removal from the matrix increased with exposure time until qeq was reached, at the state of equilibrium and for all grain sizes, on day 14. During the first 5 min, and for each grain size tested, the rate of NH4+ uptake by the zeolites was at its maximum. Similar results were reported by Taddeo et al. [25] and Kotoulas et al. [39], who found the maximum NH4+ uptake rate occurred in the first 5–10 min. qt(NH4+) for the different grain sizes diverged progressively after 40 min (Figure 1a) and decreased considerably after 3 d (or 4320 min; Figure 1b) for all grain sizes. Until the sorption equilibrium was reached for all grain sizes, the finest zeolite particle exhibited a slightly higher qt(NH4+) than the larger grain sizes. Remarkably, qt between the 1–2.5 mm and 8–16 mm grain sizes did not differ significantly (p < 0.02) beyond day 3. A steady qt was reached after 6 days for small grain sizes (1–2.5 mm and 8–16 mm), while the largest size (16–32 mm) required up to 9 days.
The NH4+ uptake rates determined can be ascribed to differences in the structures of the different zeolite grain sizes. As porous materials, zeolites are characterized by their external and internal surface areas, which interact with the surrounding solution and define the uptake rate for a given adsorptive—here, NH4+. Adsorption is defined by the bulk transport and film diffusion rates during the initial loading stages (Figure 1a), which was almost the same for all grain sizes over the first 20 min.
The finest grain size presumably favors external surface sorption due to its higher specific surface area, which explains the faster sorption rate during the first day. Similarly, the decreased rate of qt after day 6 can be explained by a reduction in the driving force and internal diffusion rate, due a reduced concentration gradient. For larger grain sizes, the greater internal diffusion rate after day 6 serves to equalize NH4+ sorption rates in the long run (Figure 1b).
Compared to the findings of Taddeo et al. [25] (stirring speed: 150 rpm; teq: 2 h; zeolite particle size: 0.2–2.0 mm), the equilibration time required in our study was quite long, especially for the 16–32 mm grain size. Nevertheless, operating at a slower stirring speed (50 rpm) was necessary to avoid zeolite grinding that would result in falsified diffusion conditions (grinding leading to a reduction in particle size and increased surface area).
From a practical point of view, the sorption kinetics of the 16–32 mm grain size seem not to be sufficiently fast to recommend the use of the largest zeolite granules tested as a single component in flow-through fixed-bed columns. However, the comparable kinetics of the 8–16 mm and 1–2.5 mm grain sizes highlights the promise of granular natural zeolites (8–16 mm) for use in drinking water treatment and the potential to reduce milling costs, by at least a factor of two, compared to finer particulate and powder-like zeolites.

3.1.2. Application of Different Reaction Kinetic Models

For a further investigation of the sorption kinetics, the experimental data shown in Figure 1 were fitted to three different reaction kinetic models (see also Figure S1). For all grain sizes, the kinetics were more accurately fit to pseudo-second order reaction kinetic plots (R2 = 1 for 1–2.5 mm, R2 = 0.999 for 8–16 mm, R2 = 0.998 for 16–32 mm). This outcome suggests that (1) the rate-limiting step is chemical sorption; and (2) the sorption rate is dependent on the sorption capacity of the zeolite and not on the concentration of NH4+ in the solution. Initially, all sorption sites are empty, and the NH4+ gradient is high. However, the zeolite becomes more and more saturated, reducing the NH4+ concentration gradient and the rate of removal (from solution). This outcome is comparable with Tang et al. [49] and Mazloomi and Jalali [50]. However, using simplified reaction kinetic models and judging their suitability based only on the correlation coefficient (R2) can lead to an overestimation of how well the data fits the model. For example, the linearization of non-linear functions can result in false correlations by reducing the error factor with log or square root transformation [45]. On the other hand, it is well known that intraparticle diffusion is typically the rate-limiting step in porous materials. Furthermore, the first- and second-order reaction kinetic models are considered empirical due to certain limitations, such as disregarding (1) the intraparticle or film diffusion that occurs; or (2) other boundary conditions. Their application is quite common for practical studies, for example, in cases (1) with adsorbents with lower porosity; (2) where film diffusion is negligible; and (3) where rather slow surface reactions occur. First- and second-order reaction kinetic models are more commonly used because they are easier to implement than more complex and precise diffusion models [44]. Due to the reasons listed above, applying exact diffusion models (film, surface, pore and intraparticle diffusion), rather than the simplified models used in many studies, is recommended [44].

3.1.3. Zeolite Sorption Capacities under Equilibrium Conditions

As demonstrated in Figure 1, qeq(NH4+) reached a state of equilibrium after 14 days. There were negligible differences in the zeolite sorption capacities with 0.130, 0.128, and 0.127 mgNH4+/gZ, corresponding to NH4+ sorption efficiencies of 98.4%, 97.8%, and 97.4%, for the 1–2.5 mm, 8–16 mm, and 16–32 mm grain sizes, respectively. Other studies observed similar sorption efficiencies for finer zeolite grain sizes [23,44,51,52,53,54]. This experiment showed that the total sorption capacity is most likely characterized by the exchangeable sites located in the zeolite’s defined inner structure and not on its external surface, hence the similar removal efficiencies. Therefore, increasing the grain size did not significantly decrease qeq(NH4+) for granular zeolites in this set-up. Furthermore, a possible temperature effect can be excluded on the basis of similar removal efficiencies being observed between 10 °C and 40 °C in Figure S2.

3.2. Application of Isothermal Models

To classify the sorption mechanisms that underlie ammonium removal by granular zeolites, our experimental equilibrium data was fitted to the most common isothermal models, Langmuir (Figure 2a) and Freundlich (Figure 2b), in Figure 2.
By comparing the linear plots of both sorption models it can be stated that the Langmuir isothermal model shows a better correlation—with an R2 value of 0.992 compared to R2 = 0.975 for the Freundlich model. A linear progression can be identified with subsequent cumulative saturation at higher concentrations.
Table 3 compares the sorption properties of zeolites from different locations used in this and other studies.
The underlying sorption mechanism cannot simply be described as solely homogeneous or heterogeneous, nor can it be ascribed entirely to chemisorption or physisorption. Several studies propose the Langmuir isothermal model for a better correlation [26,54,55,56], while others described the Freundlich model as more convenient [25,39,57,58]. However, as Lin et al. and Chen et al. described, it is most likely that ion exchange and adsorption coexist while removing NH4+ using clinoptilolite-like zeolites [55,59]. In their studies, the proportion of adsorption increased with an increase in the initial NH4+ concentration (clearly at an initial NH4+ concentration of 1000 mgN/L). According to both studies mentioned, the underlying sorption mechanism at 10 mgN/L NH4+ (the concentration that we selected as being relevant for drinking water treatment) can be attributed to ion exchange. This assumption was strengthened by balancing the initial and end point ion equivalents of NH4+, Na+, K+, Mg2+, and Ca2+ ions with an R2 = 0.997 (see Figure S3).
The varying maximum sorption capacities qmax for zeolites from different origins, shown in Table 3, are notable. Theoretically, if the number of silicon atoms (Si) replaced by aluminum atoms (Al) increases, it enhances the number of potential adsorption sites for exchangeable cations—resulting in a higher adsorption capacity. However, the opposite is actually observed when studying the Si/Al-ratios presented in Table 3. Here, replacing Si with Al leads to distortions of linkages (Si-O-Al, Si-O-Si). These distortions are the result of alterations in: (1) bond angles and lengths; (2) distances between linkage groups; (3) electronic charge strength; and (4) density distribution in the zeolite framework [60,61,62,63]. Zeolites with low Si/Al-ratios tend to possess negatively charged sites that are closer together, resulting in stronger electrostatic attraction and greater selectivity [64]. Conversely, this leads to shorter distances between hydrated NH4+ ions bonded via attractive electrostatic forces to the zeolite framework. Therefore, low Si/Al-ratios can favor electrostatic repulsion, which might explain the lower qmax at lower Si/Al-ratios observed in Table 3 [65].
These results support the findings of Section 3.1.3: the grain size of the natural zeolite under investigation does not affect the loading capacity of NH4+ ions—compare the results for larger granules to the finer zeolite grain sizes in Table 3. An increase in the initial NH4+ concentration, from 1 to 5000 mgN/L, reduced the sorption efficiency for the 8–16 mm zeolite grain size from 99% to 30% due to the limited sorption capacity (see also Figure S4). Remarkably, the sorption efficiency only declined from 99% to 95% over the range of concentrations we deemed relevant to drinking water (1–110 mgN/L NH4+), further strengthening the case for granular natural zeolites.
In conclusion to this section, three factors should be considered when applying granular natural zeolites to the treatment of drinking water: (1) where the zeolites were mined—due to the different chemical compositions of natural zeolites; (2) finding a reasonable compromise of the Si/Al-ratio to achieve optimal NH4+-removal efficiencies; and (3) the initial NH4+ concentration.

Alternative Adsorbents for Ammonium Removal

Table 4 shows a wide range of alternative adsorbents developed and used for NH4+ removal and compares the results from Section 3.2.
The different boundary conditions of (1) initial NH4+ concentration; (2) particle size; (3) adsorbent dosage; (4) chemical and physical adsorbent properties; (5) water matrix; and (6) water pH make a direct comparison difficult. Nevertheless, the used zeolite (clinoptilolite) shows comparable removal efficiencies (99% to 30%) but a lower sorption capacity (0.02–20.29 mgNH4+/gZ) than most of the demonstrated alternative adsorbents.
Other adsorbents show better sorption capacities; however, these adsorbents should meet certain criteria for their cost-effective use: (1) high removal capacity; (2) abundance; and (3) low cost.
For example, Fe3O4 nanoparticles have a far higher sorption capacity (171 mgNH4+/g) [66] than the 8–16 mm granules (20.29 mgNH4+/g) investigated in this study. However, it can be assumed that these nanoparticles are not abundant and low cost. Due to their fabrication and conditioning (chemically, thermally), the alternative adsorbents shown will not meet the criteria of (2) abundance and (3) low cost (as can be expected of the listed biosorbents).

3.3. Influence of pH

pH is an important parameter affecting (1) the net charge of an adsorbent; and (2) the dissociation equilibrium between the protonated NH4+ and its conjugated base NH3(aq). To identify the pH for maximal ESorption, the sorption of NH4+ was analyzed between pH 5 and 9 at 24 h and 14 d time points (Figure 3). The ESorption for all samples was normalized against a control containing just NH4+ (10 mgN/L) in ultrapure water without an additional buffer. To equalize the background ionic strength, NaCl was added to samples at pH 5, 6, and 7. To exclude the competing effect of Na+ ions, and to ensure that the influence of pH can explain different sorption efficiencies, a separate sample (control+NaCl) containing 10 mgN/L NH4+ with an equal amount of Na+ ions in ultrapure water without an additional buffer was used.
Maximal ESorption, at 98.7%, was observed at pH 7 in the state of equilibrium after 14 days. The samples at pH 5 and pH 6 achieved ESorption values of 97.4% and 97.3%, respectively. Further pH increases reduced the sorption efficiency: 94.9% and 92.2% at pH 8 and pH 9, respectively. Therefore, the optimal pH range for NH4+ adsorption, using the granular natural zeolite investigated, is located between pH 5 and pH 7. This range is well suited to the relevant pH range for drinking water applications. These results are also consistent with those reported in other studies [25,49,77].
The differences in ESorption can be explained by the protonation of NH4+ ions. In aqueous environments, ammonium can be found as protonated NH4+ or deprotonated NH3(aq). As a function of temperature and salinity, and at pH values above 7, the NH4+/NH3(aq)-equilibrium shifts towards NH3(aq) due to the deprotonation of NH4+. The pKa of ammonium is 9.24; at pHs above this value, the predominant species is NH3(aq) that can no longer be exchanged at the zeolites’ surface (when pH > pKa). Additionally, a significant difference in the adsorption kinetics between 24 h and 14 d, especially for pH 8 and pH 9, can be explained by the NH4+/NH3(aq)-equilibrium. Apparently, the driving force for diffusion at pH 8 and pH 9 is limited due to a higher proportion of non-exchangeable NH3(aq). A pH lower than 5 was not investigated because it can lead to the collapse, or dissolving, of natural zeolites, as reported by Zhang et al. [78], Murayama et al. [79], and Leyva-Ramos et al. [80]. Furthermore, a pH value lower than 5 or higher than 9 is not relevant for most drinking water applications.

3.3.1. Point of Zero Charge

The adsorbent-dependent point of zero charge influences the net charge of adsorbents and their adsorption behavior with regard to several cations and anions. Figure 4 shows the determination of the point of zero charge by the pH drift method between pH 2 and 9 (Figure 4a) and the mass titration method with zeolite/solution ratios of w/v = 0–11% (Figure 4b).
Both determination methods indicate that the pHPZC is located between pH 6.24 and pH 6.47. Within this pH range, the ratio between positive and negative adsorption sites is almost equal. Furthermore, the interaction between the zeolite cations or anions can be estimated. Taking this into account, the zeolite used this feature to adsorb a certain amount of anions at pH < pHPZC because the surface is weakly positive charged.
The present study demonstrated the pHPZC of a granular natural zeolite with a larger grain size for the first time. However, because the composition of natural zeolites can differ according to where they were mined from, it can be assumed that the pHPZC can also vary—and several studies contain data that supports this [81,82,83,84]. Vaičiukynienė et al., used the pH drift method to determine the pHPZC in the range of 5.0–5.4 [84]. Abatal et al., identified the pHPZC around 6.2 of a natural zeolite from Mexico [81]. The pHPZC of a Serbian (Vranska Banja) natural zeolite was in the range of pH 6.5 and 7.5 [83]. Ngeno et al. reported a pHPZC around 7.5 for a natural zeolite from Kenya [82]. Finally, the water matrix under investigation, with its specific ionic strength, can also influence the location of the pHPZC [85,86].

3.4. Competition from Cations and Anions

3.4.1. Cations

Alkali metals and alkaline earth metals are ubiquitous in surface and groundwater matrices and have the potential to compete with NH4+ ions for the sorption sites found in natural zeolites, as Figure 5 shows.
Considering the standard deviations, the lowest concentrations of competing cations (of 0.7 mmol/L) had no significant effect on the normalized qeq(NH4+). At higher concentrations, the normalized qeq(NH4+) declined, especially for sodium and most notably for potassium. The largest effect observed, in the selected concentration range, occurred for potassium at a concentration of 7 mmol/L. Such a concentration of potassium reduced the zeolite’s normalized qeq(NH4+) by 8.2%. The normalized qeq(NH4+) in the presence of Mg2+ and Ca2+ appeared to plateau at 3.5 mmol/L, with no further effects at higher concentrations. In conclusion, the sorption of NH4+ was most affected by the presence of K+, followed by Na+, Mg2+, and Ca2+. Thus, a selectivity order can be assembled as follows: K+ > Na+ > Mg2+ > Ca2+, which is generally in agreement with previous reported studies (although these can differ with respect to Mg2+ and Ca2+) [43,50].
The differing competing effects of the cations investigated can be interpreted by considering their physical characteristics and the dimensions of the framework of natural zeolites—shown in Table 5 and Table 6.
Three ring types with different channel dimensions make up the framework of natural zeolites. These Si/AlO4 ring types provide different potential sorption/exchange sites for hydrated cations. Molecular sieving effects occur, which enable cations with smaller hydrated radii to diffuse into the pores and channels of natural zeolites. These sieving effects are determined by the hydrated radii—shown in Table 6—and the channel dimensions—shown in Table 5. Horizontal ten-membered rings define one channel that Na+ and Ca2+ can occupy and Mg2+ can also be exchanged in the center of this channel. Na+ and Ca2+ can both be exchanged in the horizontal eight-membered rings. However, only K+ and NH4+, with identical hydrated radii, can diffuse into, and be exchanged in, vertical and horizontal eight-membered rings. Therefore, the dominance of K+ in the cation competition experiments can be explained by its ability to be exchanged within all ring types, much like NH4+. In addition to the fact that diffusion limits Mg2+ and Ca2+ to specific adsorption/exchangeable sites, an electrostatic repulsion effect can also occur due to their comparatively higher charge density. Furthermore, it is assumed that divalent cations, such as Mg2+ and Ca2+, can saturate potential exchange sites on the zeolites’ outer surface much faster than monovalent cations (such as K+ and Na+) can. Potential CaCO3 precipitation can be excluded in this study since the pH was clearly below the saturation limit for all molar concentrations investigated.
These findings indicate that K+ concentrations have to be considered when applying granular zeolites to the treatment of drinking water. However, as natural water matrices do not usually contain K+ concentrations as high as 7.0 mmol/L (273 mg/L), minor competing effects can generally be assumed (see also Section 3.6).

3.4.2. Anions

A range of anions can be found in surface and groundwaters, possible affecting the sorption of NH4+ by natural zeolites. In the first instance, natural zeolites are known primarily as cation exchangers. Nevertheless, in Section 3.4.1, it was shown, theoretically, that natural zeolites could adsorb anions at protonated functional groups at pHs < pHPZC. For a general confirmation the effects of NO2 (161 mg/L), NO3 (217 mg/L), PO43− (332 mg/L), and SO42− (336 mg/L), all at 3.5 mmol/L, were investigated. Note that these concentrations are not relevant to drinking water. All ESorption values were normalized against a control containing 0.71 mmol/L NH4+ in ultrapure water (Figure 6). To equalize the background ionic strength, NaCl was added to specific anion samples.
To exclude a possible influence of Na+ ions and to ensure that the influence of different anions can explain different ESorption values, a further sample (control + NaCl) was used. This contained 0.71 mmol/L NH4+ and an equal amount of Na+ ions in ultrapure water.
With regard to the normalized qeq(NH4+) values shown in Figure 6, the principal reduction of 17.4% can be attributed to the competing effect of Na+ (used as a counter ion). This effect is visible in the control+NaCl sample. Focusing on the mean values for normalized qeq(NH4+), a gradual decrease from NO2 (18.8%), to NO3 (19.8%), to HPO42−/H2PO4 (21.2%), and to SO42− (21.8%) can be observed. However, the overlapping standard deviations mean that a definite reduction order cannot be determined.
For granular natural zeolites, even the unrealistically high concentrations of the investigated anions had no significant effect on the sorption of NH4+. Drinking water sources usually contain far lower concentrations of these anions and thus their influence on the treatment of drinking water by granular natural zeolites is likely to be negligible.

3.5. Regeneration

3.5.1. Single Regeneration

The total sorption capacity of the granular natural zeolites decreased gradually by 19% over three loading steps (see Figure S5), indicating the need for regeneration to ensure the longest possible product life cycle.
Firstly, single regeneration experiments provide a basic understanding of the desorption mechanisms and enable the optimization of regeneration conditions for applications in the field. In Section 3.4.1, Na+ and especially K+ were identified as the main competitors for NH4+ sorption and therefore were selected for further regeneration studies. To this end, different molar concentrations, ranging from 1 mmol/L to 1000 mmol/L, of Na+ and K+ ions, were investigated for their ability to regenerate previously loaded zeolites (Figure 7).
It was shown that the regeneration efficiency using K+ in a single regeneration step was consistently higher than that achieved with Na+ for all molar concentrations tested. Both ions displayed an exponential increase in regeneration efficiency with increasing concentration. From a concentration of 100 mmol/L and upwards, the regeneration efficiencies diverged significantly for K+ and Na+ solutions. For example, even at a 1000 mmol/L Na+ solution, the zeolite cannot be regenerated as effectively (61.8%) as it can be using a K+ solution diluted tenfold more (72.6%). The previously loaded zeolites could only be completely regenerated using a 1000 mmol/L K+ solution. This outcome reflects the findings of Section 3.4.1 and agrees with the selectivity order and the competition for cation sorption sites in the pore and channel structures.

3.5.2. Multiple Regeneration

For applications in the field, it is necessary to conduct multiple regeneration steps in order to reuse zeolites following the application of a specific amount of NH4+ or after a predetermined operating time. To investigate multiple loading and desorption steps, zeolites were first loaded in the presence of a 0.71 mM NH4+ solution and afterward were regenerated using 100 and 1000 mmol/L K+ and Na+ solutions (Figure 8). The compositions of the exchanged cations after each loading step are shown in Figure S6.
The results shown in Figure 8 confirm the findings of Section 3.5.1 and illustrate that K+ solutions are more effective regenerative reagents than Na+ solutions. It was possible to completely regenerate the zeolites twice using a 1000 mmol/L K+ solution (d), the sorption capacity remained almost constant. It is interesting to note that a 100 mmol/L K+ solution initially regenerated the zeolite to 72.6% capacity but the second and third desorption step restored a full 100% capacity (c). A 1000 mmol/L Na+ solution could not fully regenerate the zeolites; however, the desorption efficiency increased from 61.3% to 93.5% over three regeneration cycles (b). Using a 100 mmol/L Na+ solution, the zeolites could be regenerated to 22.8% to 43.2% capacity over three cycles (a). Because an incomplete regeneration was achieved with Na+ solutions, the adsorption capacity was duly decreased.
To the best of our knowledge, this study is the first deep investigation of the sorption and desorption mechanisms of granular natural zeolites. The findings from Section 3.6 are of great interest when considering the application of granular natural zeolites to the treatment of drinking water, particularly in terms of the extended product life cycle that can be achieved by using K+ solutions to regenerate the zeolites. In most cases, the NH4+ concentration in drinking water sources is lower than for wastewater applications. This suggests that an extended loading phase can be employed and that fewer regeneration cycles will be needed. Finally, operational costs can be reduced further still by optimizing the K+ concentration necessary for complete zeolite regeneration after the initial loading step.

3.6. Influence of Natural Water Matrices

In addition to the fundamental characterization of the sorption behavior of granular natural zeolites in artificial matrices, the investigation of natural water matrices is also of great interest. Therefore, the role of different cations and dissolved organic carbon (DOC), at concentrations typically found in natural water matrices, were investigated in order to evaluate their influence on the normalized qeq(NH4+) of granular natural zeolites. Here, tap water, Elbe river water, a 1:10 mix of Elbe river water and tap water, bank filtrate, and groundwater were used as matrices.
Figure 1 shows the influence of natural water matrices on the normalized qeq(NH4+) (Figure 9a) and the initial and final DOC-concentrations after zeolite treatment (Figure 9b).
A slight reduction in the normalized qeq(NH4+) of zeolites in natural water matrices, compared to the control, was observed (Figure 9a). The normalized qeq(NH4+) decreased by ~1% in tap water and by up to ~8% in Elbe water and groundwater matrices. In the discussion of Section 3.4.1, it was noted that the normalized qeq(NH4+) was expected to decrease in natural water matrices due to the increased competition from other cations. Most notably, the relatively high K+ concentration in the groundwater matrix was expected to reduce the normalized qeq(NH4+). Table 7 and Table S1 summarizes the initial cation and DOC concentrations of the different matrices used.
The DOC concentration in the control and tap water samples was observed to increase following zeolite treatment. Conversely, the DOC in natural water matrices (with a higher initial DOC) slightly decreased following treatment (Figure 9b). The increasing DOC can be explained by residual organic matter which was released from the washed zeolites and flasks. In general, zeolites can be used to remove DOC in a targeted manner [58] and in more complex applications [92]. It is assumed that the DOC fraction has a competition effect in the present experimental set-up as Elbe water had the highest content.
As stated in Section 3.3.1, the granular natural zeolite investigated had a pHPZC between 6.24 and 6.47, indicating that the surface is mostly positively charged at pHs < pHPZC. Since the pHs used in our experiments were above the pHPZC, it can be assumed that the zeolites’ surfaces are mostly negatively charged and oppose the negative zeta potential of humic compounds in the DOC at pHs > 1.6. Nevertheless, a slight increase in the small number of positively charged sorption sites could be attributed to the removal of DOC that was observed. Furthermore, it is suspected that macromolecular humic compounds functioned to block pores, or as an additional molecular sieve (diffusion resistance) for NH4+, resulting in minor removal efficiencies [44]. However, prior autoclaving in our experiment can exclude the following factors: (1) a subsequent buildup of biofilm on the surface or in the pores; and (2) microbial conversion of NH4+.
This section showed that the normalized qeq(NH4+) of granular natural zeolites was only marginally reduced by the presence of different cations or the DOC content of natural water matrices.

4. Conclusions

This study is the first time that granular natural zeolites of different grain sizes were used in an investigation of how various process parameters influence ammonium removal efficiency. The conclusions of our study can be summarized as follows:
  • Similar sorption kinetics and capacities were exhibited by 1–2.5 mm and 8–16 mm zeolite grain sizes. For this reason, 8–16 mm granular zeolites have the potential for applications in the field, will reduce milling costs, and thus can be used as cost-effective adsorbents;
  • The Langmuir sorption model described the experimental data the most accurately, and ion exchange was revealed to be the primary mechanism behind the sorption of NH4+ at concentrations relevant to drinking water treatment;
  • A maximum sorption efficiency was identified at pH 7 and the point of zero charge was determined to be in the range of pH 6.24 to pH 6.47. A reduction in zeolite’s efficiency is not expected in the relevant pH range for drinking water applications;
  • K+ concentrations have to be considered when applying granular zeolites to the treatment of drinking water as K+ ions were the most effective competitor for NH4+ sorption to natural zeolites. Accordingly, potassium ions can be used to regenerate NH4+-loaded zeolites cost-effectively;
  • Different anions exhibited minor effects on the sorption capacity of granular zeolites and can be neglected in applications in the field;
  • The natural water matrices investigated decreased the sorption capacity of granular zeolites by up to 8%. Natural zeolites were also shown to act as adsorbents for organic matter and we showed that the DOC-concentration in a water matrix affects NH4+ sorption efficiency.
In future studies, we will validate the results reported here whilst studying sorption dynamics in a more practically relevant scenario—in-column systems. Nevertheless, this study highlights the potential of granular natural zeolites (8–16 mm) as highly-effective and cost-saving adsorbents for NH4+ removal within the field of drinking water treatment. Furthermore, besides slower sorption kinetics, no disadvantages were uncovered for the largest grain size (16–32 mm) compared to the finer zeolite grain size (1–2.5 mm) tested.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w14060939/s1, Figure S1: Fitting of experimental data of time-dependent NH4+-sorption onto natural zeolite to empirical reaction kinetic model: (a) pseudo-first order reaction kinetic model; (b) pseudo-second order reaction kinetic model; (c) intra-particle-diffusion model and (d) Elovich model; Figure S2: Influence of temperature on the sorption efficiency of granular natural zeolites; Figure S3: Initial and final ion equivalent; Figure S4: NH4+-concentration dependent sorption efficiencies of granular natural zeolites: (a) for low strengthen drinking water relevant concentrations and (b) for high strengthen wastewater conditions; Figure S5: Progress of the sorption capacity after three loading steps without regeneration; Exchanged cations after first sorption (a); second sorption (b) and third sorption (c); Table S1: General information to the investigated natural water matrices.

Author Contributions

Conceptualization, S.E.; formal analysis, S.E.; visualization, S.E.; writing—original draft preparation, S.E.; interpretation of results, S.E., S.S. and H.B.; writing—review and editing, S.E., S.S. and H.B.; support of experimental design, S.S. and H.B.; supervision, S.S. and H.B.; project administration, H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the German Federal Ministry of Education and Research (No. 02WCL1472A-I).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request.

Acknowledgments

We would like to thank HeGo Biotec GmbH for providing the zeolite used in this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jong, H.D. Living Standards in a Modernizing World—A Long-Run Perspective on Material Wellbeing and Human Development. In Global Handbook of Quality of Life; Glatzer, W., Camfield, L., Møller, V., Rojas, M., Eds.; Springer: Dordrecht, The Netherlands, 2015; pp. 45–74. ISBN 978-94-017-9177-9. [Google Scholar]
  2. Roser, M. Future Population Growth. Available online: https://ourworldindata.org/future-population-growth#global-population-growth (accessed on 11 November 2021).
  3. United Nations. World Population Prospects—Population Division—United Nations. Available online: https://population.un.org/wpp/Graphs/ (accessed on 16 July 2021).
  4. Bohnet, M.; Ullmann, F. Ullmann’s Encyclopedia of Industrial Chemistry, 6th compl. rev. ed.; Wiley-VCH: Weinheim, Germany, 2003; ISBN 9783527303854. [Google Scholar]
  5. Roser, M.; Ritchie, H. Nitrogen Fertilizers. Available online: https://ourworldindata.org/fertilizers#nitrogen-fertilizer-production (accessed on 14 November 2021).
  6. Shakya, B.M.; Nakamura, T.; Kamei, T.; Shrestha, S.D.; Nishida, K. Seasonal Groundwater Quality Status and Nitrogen Contamination in the Shallow Aquifer System of the Kathmandu Valley, Nepal. Water 2019, 11, 2184. [Google Scholar] [CrossRef] [Green Version]
  7. Zhou, Y.; Li, P.; Xue, L.; Dong, Z.; Li, D. Solute geochemistry and groundwater quality for drinking and irrigation purposes: A case study in Xinle City, North China. Geochemistry 2020, 80, 125609. [Google Scholar] [CrossRef]
  8. Rusydi, A.F. Estimation of ammonium sources in Indonesian coastal alluvial groundwater using Cl and GIS. GEOMATE 2019, 17, 53–58. [Google Scholar] [CrossRef]
  9. Rusydi, A.F.; Onodera, S.-I.; Saito, M.; Hyodo, F.; Maeda, M.; Sugianti, K.; Wibawa, S. Potential Sources of Ammonium-Nitrogen in the Coastal Groundwater Determined from a Combined Analysis of Nitrogen Isotope, Biological and Geological Parameters, and Land Use. Water 2021, 13, 25. [Google Scholar] [CrossRef]
  10. Fitts, C.R. Groundwater Science, 2nd ed.; Academic: Oxford, UK, 2013; ISBN 0123847052. [Google Scholar]
  11. Elshorbagy, W. (Ed.) Water Treatment; InTech: London, UK, 2013; ISBN 978-953-51-0928-0. [Google Scholar]
  12. McCarty, P.L. What is the Best Biological Process for Nitrogen Removal: When and Why? Environ. Sci. Technol. 2018, 52, 3835–3841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wei, W.; Zhou, X.; Wang, D.; Sun, J.; Wang, Q. Free ammonia pre-treatment of secondary sludge significantly increases anaerobic methane production. Water Res. 2017, 118, 12–19. [Google Scholar] [CrossRef]
  14. Carrera, J.; Baeza, J.A.; Vicent, T.; Lafuente, J. Biological nitrogen removal of high-strength ammonium industrial wastewater with two-sludge system. Water Res. 2003, 37, 4211–4221. [Google Scholar] [CrossRef]
  15. WHO. Guidelines for Drinking-Water Quality: Fourth Edition Incorporating the First Addendum; World Health Organisation: Geneva, Switzerland, 2017. [Google Scholar]
  16. Du, Y.; Ma, T.; Deng, Y.; Shen, S.; Lu, Z. Sources and fate of high levels of ammonium in surface water and shallow groundwater of the Jianghan Plain, Central China. Environ. Sci. Process. Impacts 2017, 19, 161–172. [Google Scholar] [CrossRef] [PubMed]
  17. de Vet, W.W.J.M.; van Genuchten, C.C.A.; van Loosdrecht, M.C.M.; van Dijk, J.C. Water quality and treatment of river bank filtrate. Drink. Water Eng. Sci. 2010, 3, 79–90. [Google Scholar] [CrossRef] [Green Version]
  18. Berg, M.; Tran, H.C.; Nguyen, T.C.; Pham, H.V.; Schertenleib, R.; Giger, W. Arsenic contamination of groundwater and drinking water in Vietnam: A human health threat. Environ. Sci. Technol. 2001, 35, 2621–2626. [Google Scholar] [CrossRef] [Green Version]
  19. Norrman, J.; Sparrenbom, C.J.; Berg, M.; Dang, D.N.; Jacks, G.; Harms-Ringdahl, P.; Pham, Q.N.; Rosqvist, H. Tracing sources of ammonium in reducing groundwater in a well field in Hanoi (Vietnam) by means of stable nitrogen isotope (δ15N) values. Appl. Geochem. 2015, 61, 248–258. [Google Scholar] [CrossRef]
  20. Jiao, J.J.; Wang, Y.; Cherry, J.A.; Wang, X.; Zhi, B.; Du, H.; Wen, D. Abnormally high ammonium of natural origin in a coastal aquifer-aquitard system in the Pearl River Delta, China. Environ. Sci. Technol. 2010, 44, 7470–7475. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, Z.; Xiao, C.; Adeyeye, O.; Yang, W.; Liang, X. Source and Mobilization Mechanism of Iron, Manganese and Arsenic in Groundwater of Shuangliao City, Northeast China. Water 2020, 12, 534. [Google Scholar] [CrossRef] [Green Version]
  22. Verordnung über die Qualität von Wasser für den Menschlichen Gebrauch (Trinkwasserverordnung—TrinkwV). Available online: http://www.gesetze-im-internet.de/trinkwv_2001/BJNR095910001.html#BJNR095910001BJNG000201310 (accessed on 3 August 2021).
  23. Gupta, V.K.; Sadegh, H.; Yari, M.; Shahryari Ghoshekandi, R.; Maazinejad, B.; Chahardori, M. Removal of ammonium ions from wastewater: A short review in development of efficient methods. Glob. J. Environ. Sci. Manag. 2015, 1, 149–158. [Google Scholar] [CrossRef]
  24. Uhlmann, W.; Kreuziger, Y.; Kruspe, R.; Neumann, J.; Berg, T. Untersuchungen zu Ammonium in Ostsächsischen Bergbaufolgeseen. Available online: https://www.wasser.sachsen.de/download/Abschlussbericht_TP_05_online.pdf (accessed on 4 August 2021).
  25. Taddeo, R.; Prajapati, S.; Lepistö, R. Optimizing ammonium adsorption on natural zeolite for wastewaters with high loads of ammonium and solids. J. Porous Mater. 2017, 24, 1545–1554. [Google Scholar] [CrossRef]
  26. Guaya, D.; Valderrama, C.; Farran, A.; Armijos, C.; Cortina, J.L. Simultaneous phosphate and ammonium removal from aqueous solution by a hydrated aluminum oxide modified natural zeolite. Chem. Eng. J. 2015, 271, 204–213. [Google Scholar] [CrossRef] [Green Version]
  27. Karadag, D.; Tok, S.; Akgul, E.; Turan, M.; Ozturk, M.; Demir, A. Ammonium removal from sanitary landfill leachate using natural Gördes clinoptilolite. J. Hazardous Mater. 2008, 153, 60–66. [Google Scholar] [CrossRef] [PubMed]
  28. Du, Q.; Liu, S.; Cao, Z.; Wang, Y. Ammonia removal from aqueous solution using natural Chinese clinoptilolite. Sep. Purif. Technol. 2005, 44, 229–234. [Google Scholar] [CrossRef]
  29. U.S. Geological Survey. Mineral Commodity Summaries 2021. Available online: https://pubs.usgs.gov/periodicals/mcs2021/mcs2021.pdf (accessed on 2 October 2021).
  30. Virta, R. Minerals Yearbook 2011. Available online: https://s3-us-west-2.amazonaws.com/prd-wret/assets/palladium/production/mineral-pubs/zeolites/myb1-2011-zeoli.pdf (accessed on 12 December 2021).
  31. Zeolite Products. Home—Zeolite Products. Available online: https://www.zeolite-products.com/ (accessed on 15 March 2022).
  32. Wang, S.; Peng, Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chem. Eng. J. 2010, 156, 11–24. [Google Scholar] [CrossRef]
  33. Morante-Carballo, F.; Montalván-Burbano, N.; Carrión-Mero, P.; Jácome-Francis, K. Worldwide Research Analysis on Natural Zeolites as Environmental Remediation Materials. Sustainability 2021, 13, 6378. [Google Scholar] [CrossRef]
  34. Huang, J.; Kankanamge, N.R.; Chow, C.; Welsh, D.T.; Li, T.; Teasdale, P.R. Removing ammonium from water and wastewater using cost-effective adsorbents: A review. J. Environ. Sci. 2018, 63, 174–197. [Google Scholar] [CrossRef] [PubMed]
  35. Weatherley, L.R.; Miladinovic, N.D. Comparison of the ion exchange uptake of ammonium ion onto New Zealand clinoptilolite and mordenite. Water Res. 2004, 38, 4305–4312. [Google Scholar] [CrossRef] [PubMed]
  36. Erdem, E.; Karapinar, N.; Donat, R. The removal of heavy metal cations by natural zeolites. J. Colloid Interface Sci. 2004, 280, 309–314. [Google Scholar] [CrossRef]
  37. Belviso, C. Zeolite for Potential Toxic Metal Uptake from Contaminated Soil: A Brief Review. Processes 2020, 8, 820. [Google Scholar] [CrossRef]
  38. Kwakye-Awuah, B.; Sefa-Ntiri, B.; Von-Kiti, E.; Nkrumah, I.; Williams, C. Adsorptive Removal of Iron and Manganese from Groundwater Samples in Ghana by Zeolite Y Synthesized from Bauxite and Kaolin. Water 2019, 11, 1912. [Google Scholar] [CrossRef] [Green Version]
  39. Kotoulas, A.; Agathou, D.; Triantaphyllidou, I.; Tatoulis, T.; Akratos, C.; Tekerlekopoulou, A.; Vayenas, D. Zeolite as a Potential Medium for Ammonium Recovery and Second Cheese Whey Treatment. Water 2019, 11, 136. [Google Scholar] [CrossRef] [Green Version]
  40. Collison, R.; Grismer, M. Upscaling the Zeolite-Anammox Process: Treatment of Anaerobic Digester Filtrate. Water 2018, 10, 1553. [Google Scholar] [CrossRef] [Green Version]
  41. Collison, R.S.; Grismer, M.E. Upscaling the Zeolite-Anammox Process: Treatment of Secondary Effluent. Water 2018, 10, 236. [Google Scholar] [CrossRef] [Green Version]
  42. Yapsakli, K.; Aktan, C.K.; Mertoglu, B. Anammox-zeolite system acting as buffer to achieve stable effluent nitrogen values. Biodegradation 2017, 28, 69–79. [Google Scholar] [CrossRef]
  43. Kodewitz, P. Datasheet CLP85+ Natural Zeolite; Zeolite: Berlin, Germany; Available online: https://www.zeolith-gmbh.com/produkte/klinoptilolith-zeolith (accessed on 16 February 2022).
  44. Worch, E. Adsorption Technology in Water Treatment: Fundamentals, Processes, and Modeling, 2nd ed.; De Gruyter: Berlin, Germany, 2021; ISBN 9783110715507. [Google Scholar]
  45. William Kajjumba, G.; Emik, S.; Öngen, A.; Kurtulus Özcan, H.; Aydın, S. (Eds.) Modelling of Adsorption Kinetic Processes—Errors, Theory and Application; IntechOpen: London, UK, 2019; ISBN 978-1-78984-819-9. [Google Scholar]
  46. Edebali, S. (Ed.) Advanced Sorption Process Applications; IntechOpen: London, UK, 2019; ISBN 978-1-78984-818-2. [Google Scholar]
  47. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J 2010, 156, 2–10. [Google Scholar] [CrossRef]
  48. Fiol, N.; Villaescusa, I. Determination of sorbent point zero charge: Usefulness in sorption studies. Environ. Chem. Lett. 2009, 7, 79–84. [Google Scholar] [CrossRef]
  49. Tang, H.; Xu, X.; Wang, B.; Lv, C.; Shi, D. Removal of Ammonium from Swine Wastewater Using Synthesized Zeolite from Fly Ash. Sustainability 2020, 12, 3423. [Google Scholar] [CrossRef] [Green Version]
  50. Mazloomi, F.; Jalali, M. Ammonium removal from aqueous solutions by natural Iranian zeolite in the presence of organic acids, cations and anions. J. Environ. Chem. Eng. 2016, 4, 1664–1673. [Google Scholar] [CrossRef]
  51. Singh, A.K.; Bhagowati, S.; Das, T.K.; Yubbe, D.; Rahmann, B.; Nath, M.; Obing, P.; Singh, W.S.K.; Renthlei, C.Z.; Pachuau, L.; et al. Thakur. Assessment of Arsenic, Fluoride, Iron, Nitrate and Heavy Metals in Drinking Water of Northeastern India. Himal. Ecol. 2008, 1–7. [Google Scholar]
  52. Schierano, M.C.; Maine, M.A.; Panigatti, M.C. Dairy farm wastewater treatment using horizontal subsurface flow wetlands with Typha domingensis and different substrates. Environ. Technol. 2017, 38, 192–198. [Google Scholar] [CrossRef] [PubMed]
  53. Stefanakis, A.I.; Tsihrintzis, V.A. Use of zeolite and bauxite as filter media treating the effluent of Vertical Flow Constructed Wetlands. Microporous Mesoporous Mater. 2012, 155, 106–116. [Google Scholar] [CrossRef]
  54. Huo, H.; Lin, H.; Dong, Y.; Cheng, H.; Wang, H.; Cao, L. Ammonia-nitrogen and phosphates sorption from simulated reclaimed waters by modified clinoptilolite. J. Hazard. Mater. 2012, 229-230, 292–297. [Google Scholar] [CrossRef]
  55. Lin, L.; Lei, Z.; Wang, L.; Liu, X.; Zhang, Y.; Wan, C.; Lee, D.-J.; Tay, J.H. Adsorption mechanisms of high-levels of ammonium onto natural and NaCl-modified zeolites. Sep. Purif. Technol. 2013, 103, 15–20. [Google Scholar] [CrossRef] [Green Version]
  56. Ivanova, E.; Karsheva, M.; Koumanova, B. Adsorption of ammonium ions onto natural zeolite. J. Univ. Chem. Technol. Metall. 2010, 45, 295–301. [Google Scholar]
  57. Cheng, Z.; Ding, W. Ammonium removal from water by natural and modified zeolite: Kinetic, equilibrium, and thermodynamic studies. Desalin. Water Treat. 2015, 55, 978–985. [Google Scholar] [CrossRef]
  58. Moussavi, G.; Talebi, S.; Farrokhi, M.; Sabouti, R.M. The investigation of mechanism, kinetic and isotherm of ammonia and humic acid co-adsorption onto natural zeolite. Chem. Eng. J. 2011, 171, 1159–1169. [Google Scholar] [CrossRef]
  59. Chen, H.-F.; Lin, Y.-J.; Chen, B.-H.; Yoshiyuki, I.; Liou, S.; Huang, R.-T. A Further Investigation of NH4+ Removal Mechanisms by Using Natural and Synthetic Zeolites in Different Concentrations and Temperatures. Minerals 2018, 8, 499. [Google Scholar] [CrossRef] [Green Version]
  60. Senchenya, I.N.; Kazanskii, V.B.; Beran, S. Quantum chemical study of the effect of the structural characteristics of zeolites on the properties of their bridging hydroxyl groups. Part 2. J. Phys. Chem. 1968, 90, 4857–4859. [Google Scholar] [CrossRef]
  61. Katada, N.; Suzuki, K.; Noda, T.; Sastre, G.; Niwa, M. Correlation between Brønsted Acid Strength and Local Structure in Zeolites. J. Phys. Chem. C 2009, 113, 19208–19217. [Google Scholar] [CrossRef]
  62. Noda, T.; Suzuki, K.; Katada, N.; Niwa, M. Combined study of IRMS-TPD measurement and DFT calculation on Brønsted acidity and catalytic cracking activity of cation-exchanged Y zeolites. J. Catal. 2008, 259, 203–210. [Google Scholar] [CrossRef]
  63. Sastre, G.; Katada, N.; Niwa, M. Computational Study of Brønsted Acidity of Mordenite. Effect of the Electric Field on the Infrared OH Stretching Frequencies. J. Phys. Chem. C 2010, 114, 15424–15431. [Google Scholar] [CrossRef]
  64. Herron, N.; Corbin, D.R. (Eds.) Inclusion Chemistry with Zeolites: Nanoscale Materials by Design; Springer: Dordrecht, The Netherlands, 1995; ISBN 9789401101196. [Google Scholar]
  65. Van Santen, R.A. Molecular Heterogeneous Catalysis: A Conceptual and Computational Approach; Wiley-VCH: Weinheim, Germany, 2006; ISBN 3527610847. [Google Scholar]
  66. Zare, K.; Sadegh, H.; Shahryari-ghoshekandi, R.; Asif, M.; Tyagi, I.; Agarwal, S.; Gupta, V.K. Equilibrium and kinetic study of ammonium ion adsorption by Fe3O4 nanoparticles from aqueous solutions. J. Mol. Liq. 2016, 213, 345–350. [Google Scholar] [CrossRef]
  67. Thornton, A.; Pearce, P.; Parsons, S.A. Ammonium removal from digested sludge liquors using ion exchange. Water Res. 2007, 41, 433–439. [Google Scholar] [CrossRef]
  68. Otal, E.; Vilches, L.F.; Luna, Y.; Poblete, R.; García-Maya, J.M.; Fernández-Pereira, C. Ammonium Ion Adsorption and Settleability Improvement Achieved in a Synthetic Zeolite-Amended Activated Sludge. Chin. J. Chem. Eng. 2013, 21, 1062–1068. [Google Scholar] [CrossRef]
  69. Sica, M.; Duta, A.; Teodosiu, C.; Draghici, C. Thermodynamic and kinetic study on ammonium removal from a synthetic water solution using ion exchange resin. Clean. Technol. Environ. Policy 2014, 16, 351–359. [Google Scholar] [CrossRef]
  70. Jorgensen, T.C.; Weatherley, L.R. Ammonia removal from wastewater by ion exchange in the presence of organic contaminants. Water Res. 2003, 37, 1723–1728. [Google Scholar] [CrossRef]
  71. Liao, Z.-L.; Chen, H.; Zhu, B.-R.; Li, H.-Z. Combination of powdered activated carbon and powdered zeolite for enhancing ammonium removal in micro-polluted raw water. Chemosphere 2015, 134, 127–132. [Google Scholar] [CrossRef]
  72. Kizito, S.; Wu, S.; Kipkemoi Kirui, W.; Lei, M.; Lu, Q.; Bah, H.; Dong, R. Evaluation of slow pyrolyzed wood and rice husks biochar for adsorption of ammonium nitrogen from piggery manure anaerobic digestate slurry. Sci. Total Environ. 2015, 505, 102–112. [Google Scholar] [CrossRef] [PubMed]
  73. Juan, R.; Hernández, S.; Andrés, J.M.; Ruiz, C. Ion exchange uptake of ammonium in wastewater from a Sewage Treatment Plant by zeolitic materials from fly ash. J. Hazard. Mater. 2009, 161, 781–786. [Google Scholar] [CrossRef] [PubMed]
  74. Zhao, Y.; Niu, Y.; Hu, X.; Xi, B.; Peng, X.; Liu, W.; Guan, W.; Wang, L. Removal of ammonium ions from aqueous solutions using zeolite synthesized from red mud. Desalin. Water Treat. 2015, 57, 4720–4731. [Google Scholar] [CrossRef]
  75. Yu, X.; Wan, C.; Lei, Z.; Liu, X.; Zhang, Y.; Lee, D.-J.; Tay, J.-H. Adsorption of ammonium by aerobic granules under high ammonium levels. J. Taiwan Inst. Chem. Eng. 2014, 45, 202–206. [Google Scholar] [CrossRef]
  76. Ismail, Z.Z.; Tezel, U.; Pavlostathis, S.G. Sorption of quaternary ammonium compounds to municipal sludge. Water Res. 2010, 44, 2303–2313. [Google Scholar] [CrossRef] [PubMed]
  77. Wasielewski, S.; Rott, E.; Minke, R.; Steinmetz, H. Evaluation of Different Clinoptilolite Zeolites as Adsorbent for Ammonium Removal from Highly Concentrated Synthetic Wastewater. Water 2018, 10, 584. [Google Scholar] [CrossRef] [Green Version]
  78. Zhang, M.; Zhang, H.; Xu, D.; Han, L.; Niu, D.; Tian, B.; Zhang, J.; Zhang, L.; Wu, W. Removal of ammonium from aqueous solutions using zeolite synthesized from fly ash by a fusion method. Desalination 2011, 271, 111–121. [Google Scholar] [CrossRef]
  79. Murayama, N.; Yoshida, S.; Takami, Y.; Yamamoto, H.; Shibata, J. Simultaneous Removal of NH4+ and PO43− in Aqueous Solution and Its Mechanism by Using Zeolite Synthesized from Coal Fly Ash. Sep. Sci. Technol. 2003, 38, 113–130. [Google Scholar] [CrossRef]
  80. Leyva-Ramos, R.; Aguilar-Armenta, G.; Gonzalez-Gutierrez, L.V.; Guerrero-Coronado, R.M.; Mendoza-Barron, J. Ammonia exchange on clinoptilolite from mineral deposits located in Mexico. J. Chem. Technol. Biotechnol. 2004, 79, 651–657. [Google Scholar] [CrossRef]
  81. Abatal, M.; Córdova Quiroz, A.V.; Olguín, M.T.; Vázquez-Olmos, A.R.; Vargas, J.; Anguebes-Franseschi, F.; Giácoman-Vallejos, G. Sorption of Pb(II) from Aqueous Solutions by Acid-Modified Clinoptilolite-Rich Tuffs with Different Si/Al Ratios. Appl. Sci. 2019, 9, 2415. [Google Scholar] [CrossRef] [Green Version]
  82. Ngeno, E.C.; Shikuku, V.O.; Orata, F.; Baraza, L.D.; Kimosop, S.J. Caffeine and Ciprofloxacin Adsorption from Water onto Clinoptilolite: Linear Isotherms, Kinetics, Thermodynamic and Mechanistic Studies. S. Afr. J. Chem. 2019, 72, 136–142. [Google Scholar] [CrossRef]
  83. Sljivic, M.; Smiciklas, I.; Pejanovic, S.; Plecas, I. Comparative study of Cu2+ adsorption on a zeolite, a clay and a diatomite from Serbia. Appl. Clay Sci. 2009, 43, 33–40. [Google Scholar] [CrossRef]
  84. Vaičiukynienė, D.; Mikelionienė, A.; Baltušnikas, A.; Kantautas, A.; Radzevičius, A. Removal of ammonium ion from aqueous solutions by using unmodified and H2O2-modified zeolitic waste. Sci. Rep. 2020, 10, 352. [Google Scholar] [CrossRef] [PubMed]
  85. Mahmood, T.; Saddique, M.T.; Naeem, A.; Westerhoff, P.; Mustafa, S.; Alum, A. Comparison of Different Methods for the Point of Zero Charge Determination of NiO. Ind. Eng. Chem. Res. 2011, 50, 10017–10023. [Google Scholar] [CrossRef]
  86. Liu, X.; Mäki-Arvela, P.; Aho, A.; Vajglova, Z.; Gun’ko, V.M.; Heinmaa, I.; Kumar, N.; Eränen, K.; Salmi, T.; Murzin, D.Y. Zeta Potential of Beta Zeolites: Influence of Structure, Acidity, pH, Temperature and Concentration. Molecules 2018, 23, 946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Inglezakis, V.J.; Zorpas, A. (Eds.) Handbook of Natural Zeolites: Natural Zeolites Structure and Porosity; Bentham Science Publishers: Sharjah, United Arab Emirates, 2012; ISBN 978-1-60805-446-6. [Google Scholar]
  88. Margeta, K.; Zabukovec, N.; Siljeg, M.; Farkas, A. Natural Zeolites in Water Treatment—How Effective is Their Use. In Water Treatment; Elshorbagy, W., Ed.; InTech: London, UK, 2013; ISBN 978-953-51-0928-0. [Google Scholar]
  89. Sidey, V. On the effective ionic radii for ammonium. Acta Crystallogr. B Struct. Sci. Cryst. Eng. Mater. 2016, 72, 626–633. [Google Scholar] [CrossRef]
  90. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  91. Nightingale, E.R., Jr. Phenomenological Theory of Ion Solvation. Effective Radii of Hydrated Ions. J. Phys. Chem. 1959, 63, 1381–1387. [Google Scholar] [CrossRef]
  92. Akcay, M.U.; Avdan, Z.Y.; Inan, H. Effect of biofiltration process on the control of THMs and HAAs in drinking water. Desalin. Water Treat. 2016, 57, 2546–2554. [Google Scholar] [CrossRef]
Figure 1. Time-dependent ammonium sorption capacity qt of CLP85+ with different grain sizes as a function of the exposure time (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(NH4+) = 10 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 1–2.5 mm, 8–16 mm, 16–32 mm; n = 3).
Figure 1. Time-dependent ammonium sorption capacity qt of CLP85+ with different grain sizes as a function of the exposure time (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(NH4+) = 10 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 1–2.5 mm, 8–16 mm, 16–32 mm; n = 3).
Water 14 00939 g001
Figure 2. Linear plots of Langmuir (a) and Freundlich (b) sorption models for NH4+ sorption on natural zeolite for the full concentration range investigated (matrix: ultrapure water; teq: up to 90 d; pH0: 5.1–5.9; c0(NH4+) = 1–5000 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 2. Linear plots of Langmuir (a) and Freundlich (b) sorption models for NH4+ sorption on natural zeolite for the full concentration range investigated (matrix: ultrapure water; teq: up to 90 d; pH0: 5.1–5.9; c0(NH4+) = 1–5000 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g002
Figure 3. Influence of pH on the normalized qeq(NH4+) after 24 h and 14 d (matrix: ultrapure water; teq: 14 d; pHcontrol = 5.9; pHcontrol+NaCl: 6.2; c0(NH4+) = 10 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 3. Influence of pH on the normalized qeq(NH4+) after 24 h and 14 d (matrix: ultrapure water; teq: 14 d; pHcontrol = 5.9; pHcontrol+NaCl: 6.2; c0(NH4+) = 10 mgN/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g003
Figure 4. Point of zero charge determination: pH drift method (a,b) and mass titration method (c) (matrix: ultrapure water; teq: 1 d; c(NaCl) = 0.03 M; pH0: 2–9; w/v: 0–11%; T = 22 °C; grain size: 8–16 mm).
Figure 4. Point of zero charge determination: pH drift method (a,b) and mass titration method (c) (matrix: ultrapure water; teq: 1 d; c(NaCl) = 0.03 M; pH0: 2–9; w/v: 0–11%; T = 22 °C; grain size: 8–16 mm).
Water 14 00939 g004
Figure 5. Normalized qeq(NH4+) in the presence of competing cations in three different molar concentrations (matrix: ultrapure water; teq: 14 d; pHcontrol: 5.9; pH0 = 6.2; c0(NH4+) = 0.71 mmol/L; c0(cations) = 0.7, 3.5, 7.0 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 5. Normalized qeq(NH4+) in the presence of competing cations in three different molar concentrations (matrix: ultrapure water; teq: 14 d; pHcontrol: 5.9; pH0 = 6.2; c0(NH4+) = 0.71 mmol/L; c0(cations) = 0.7, 3.5, 7.0 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g005
Figure 6. Normalized qeq(NH4+) in the presence of competing anions and Na+ (matrix: ultrapure water; teq: 14 d; pHcontrol: 5.9; pHcontrol+NaCl: 6.2; pH0: 4.9–6.2; c0(NH4+) = 0.71 mmol/L; c0(anions) = 3.5 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 6. Normalized qeq(NH4+) in the presence of competing anions and Na+ (matrix: ultrapure water; teq: 14 d; pHcontrol: 5.9; pHcontrol+NaCl: 6.2; pH0: 4.9–6.2; c0(NH4+) = 0.71 mmol/L; c0(anions) = 3.5 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g006
Figure 7. Regeneration efficiencies using Na+ and K+ ions at different molar concentrations following a single sorption step (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(Na+, K+) = 1–1000 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 7. Regeneration efficiencies using Na+ and K+ ions at different molar concentrations following a single sorption step (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(Na+, K+) = 1–1000 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g007
Figure 8. Sorption capacities and regeneration efficiencies during multiple rounds of NH4+ loading and regeneration (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(NH4+) = 0.71 mmol/L; c0(Na+, K+) = 100, 1000 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 8. Sorption capacities and regeneration efficiencies during multiple rounds of NH4+ loading and regeneration (matrix: ultrapure water; teq: 14 d; pH0: 5.9; c0(NH4+) = 0.71 mmol/L; c0(Na+, K+) = 100, 1000 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g008
Figure 9. Influence of different natural water matrices on the normalized qeq(NH4+) of a natural zeolite (a); initial and final DOC-concentrations after zeolite treatment (b) (matrix: natural; teq: 14 d; pHcontrol: 5.9; pH0: 7.8; c0(NH4+) = 0.71 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Figure 9. Influence of different natural water matrices on the normalized qeq(NH4+) of a natural zeolite (a); initial and final DOC-concentrations after zeolite treatment (b) (matrix: natural; teq: 14 d; pHcontrol: 5.9; pH0: 7.8; c0(NH4+) = 0.71 mmol/L; w/v = 100 gZ/L; T = 22 °C; grain size: 8–16 mm; n = 3).
Water 14 00939 g009
Table 1. Chemical composition and zeolite characteristics, data provided by supplier [43].
Table 1. Chemical composition and zeolite characteristics, data provided by supplier [43].
CompositionValue (%)Characteristics
SiO265.00–71.30Exchange capacity1.2–1.5 mol/kg
Al2O311.50–13.10SelectivityNH4+ > K+ > Na+ > Ca2+ > Mg2+
CaO2.70–5.20Mean pore diameter0.4 nm
K2O2.20–3.40Specific surface30–60 m2/g
Fe2O30.70–1.90Si/Al4.80–5.40 (−)
MgO0.60–1.20Grain sizes1–2.5 mm
Na2O0.20–1.308–16 mm
TiO20.10–0.3016–32 mm
Table 2. Composition of the used 0.1 M buffer systems.
Table 2. Composition of the used 0.1 M buffer systems.
BufferChemical AChemical BA (mol/L)B (mol/L)
pH 5C2H4O2C2H3NaO20.030.07
pH 6Na2HPO4NaH2PO4·H2O0.010.09
pH 7Na2HPO4NaH2PO4·H2O0.060.04
pH 8Na2B4O7·10H2OHCl0.1x
pH 9NaH2PO4·H2OHCl0.1x
X adjusted by 0.1 M HCl.
Table 3. Experimental isotherm constants compared to other studies.
Table 3. Experimental isotherm constants compared to other studies.
ZeoliteGrain Size (mm)Si/Al (−)c0(NH4+) (mg/L)LangmuirFreundlich
KL (L/mg)qmax (mg/L)R2 (−)KF (L/g)n (−)R2 (−)
CLP85+ (a)8–164.8–5.41–68900.00421.30.9921.210.330.975
Slovakia (b)<0.25.610–50000.006330.991.840.360.97
China (c)0.8–1.433.3810–40000.00914.30.9930.980.360.973
(a) this study, (b) [26], (c) [55].
Table 4. Alternative adsorbents used for NH4+ removal, summarized from Huang et al. [34] and Gupta et al. [23].
Table 4. Alternative adsorbents used for NH4+ removal, summarized from Huang et al. [34] and Gupta et al. [23].
Alternative
Adsorbent
Product NameGrain Size (mm)c0(NH4+) (mg/L)Capacity (mgNH4+/g)Removal
Efficiency (%)
Reference
ClinoptiloliteCLP85+8–162–68900.02–20.2999–30This study
Other zeolitesMesoliteNA x505595[67]
Synthetic zeoliteNA x181370[68]
Polymeric ion exchangerPurolite C 150 HNA x32–19336>65[69]
Dowex 50w-x8 and Purolite MN500~0.1832–25751NA x[70]
Carbon-based adsorbentsCombination of PAC xx and PZ xxx0.50–0.8523NAx49[71]
Rice husk biochar0.25–0.50322–18005184[72]
Industrial wastesFly ash (thermally activated)<0.8067784[73]
Red mud~0.156–64422>50[74]
NanoparticlesFe3O4NA x103–180171NA x[66]
BiosorbentsAerobic granulesNA x30024NA x[75]
Municipal sludgeNA x300NA x89[76]
x NA: not available; xx PAC: powdered activated carbon; xxx PZ: powdered zeolite.
Table 5. Channel dimensions and ring types of natural zeolites.
Table 5. Channel dimensions and ring types of natural zeolites.
Channel Dimension (nm)Ring Type
Inglezakis and Zorpas [87]Margeta et al. [88]NumberArrangement
0.75·0.310.72·0.4410Horizontal
0.47·0.280.55·0.408Vertical
0.46·0.360.47·0.418Horizontal
Table 6. Physical characteristics of the investigated cations.
Table 6. Physical characteristics of the investigated cations.
CationCharge Density (C/mm3)Ionic Radius (nm)Hydrated Radius (nm)
NH4+120.148 (a)0.331 (c)
K+140.138 (b)0.331 (c)
Na+370.102 (b)0.358 (c)
Ca2+760.100 (b)0.412 (c)
Mg2+2050.072 (b)0.428 (c)
(a) [89], (b) [90], (c) [91].
Table 7. Initial cation and DOC concentrations of the investigated natural water matrices.
Table 7. Initial cation and DOC concentrations of the investigated natural water matrices.
Water MatricesCations (mmol/L)DOC(mg/L)
K+Na+Mg2+Ca2+
Ulrapure water (control)<CR x<CR x<CR x<CR x0.09
Tap water0.040.380.300.411.77
Bank filtrate0.250.911.720.781.77
Elbe-/tap water (1:10)0.050.460.360.442.39
Groundwater1.751.171.130.915.00
Elbe river water0.110.150.940.566.59
x below calibration range.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Eberle, S.; Börnick, H.; Stolte, S. Granular Natural Zeolites: Cost-Effective Adsorbents for the Removal of Ammonium from Drinking Water. Water 2022, 14, 939. https://doi.org/10.3390/w14060939

AMA Style

Eberle S, Börnick H, Stolte S. Granular Natural Zeolites: Cost-Effective Adsorbents for the Removal of Ammonium from Drinking Water. Water. 2022; 14(6):939. https://doi.org/10.3390/w14060939

Chicago/Turabian Style

Eberle, Stephan, Hilmar Börnick, and Stefan Stolte. 2022. "Granular Natural Zeolites: Cost-Effective Adsorbents for the Removal of Ammonium from Drinking Water" Water 14, no. 6: 939. https://doi.org/10.3390/w14060939

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop