Fouling behaviour in membrane distillation (MD) processes plays a crucial role in determining their widespread acceptability. Most studies have primarily focused on model organic foulants, such as humic acid (HA) and sodium alginate (SA). This study investigates the fouling of a polytetrafluoroethylene membrane in a direct contact MD (DCMD) using model organics (i.e., HA and SA) and real wastewater. The results indicated that the flux decline (5–60%) was only observed during the initial phase of the operation with model organic foulants. In contrast, real wastewater caused a gradual decline in flux throughout the experiment in both the concentrate (40%) and continuous (90%) modes. The study also found significant differences in the fouling layer morphology, composition, and hydrophobicity between the model organic foulants and real wastewater. Fourier transform infrared spectroscopy findings demonstrated that the fouling layer formed by real wastewater varied significantly from model organics, which primarily comprised of protein-like and polysaccharide-like functional groups. Finally, liquid chromatography–organic carbon detection revealed that the fouling layer of the MD membrane with real wastewater was composed of 40.7% hydrophobic and 59.3% hydrophilic organics. This study suggests that model organics may not accurately reflect real wastewater fouling.

  • Model organics and real wastewater exhibit different fouling behaviours in membrane distillation (MD).

  • Fouling layer by model organics was unevenly distributed and affected MD membrane properties.

  • Long-term (22 days) experiment showed continuous flux decline caused by organic–inorganic foulant interaction.

  • Real wastewater treatment instigated a membrane fouling layer consisting of proteins and polysaccharides.

Erratic rainfall patterns and associated runoff owing to climate change are one of the main causes of water scarcity, resulting in an uneven distribution of water resources worldwide. This necessitates the development of simple and holistic solutions to ensure a continuous and reliable supply of freshwater (Manikandan et al. 2022). In areas where potable water is in short or limited supply, the treated effluent (TE) generated by wastewater treatment plants (WWTPs) can be used as a viable alternative (Zhang & Liu 2021). Owing to its reduced sensitivity to weather and seasonal fluctuations, reclaimed water, which is an underutilised resource, should be considered for potable and non-potable water reuse applications (Machineni 2020). To meet the growing demand for clean water and reduce the associated environmental implications, state-of-the-art membrane-based separation processes such as reverse osmosis (RO) and nanofiltration (NF) have already been applied (Al-Araji et al. 2022; Abbas et al. 2023). Notably, in recent years, increased focus has been placed on membrane distillation (MD) technology for advanced wastewater treatment (Ramlow et al. 2017; Liu et al. 2020).

MD is a thermal technique that operates based on the principle of pressure difference between the hot feed and cold permeate separated by a microporous hydrophobic membrane (Ramezanianpour & Sivakumar 2014a). It is theoretically possible to obtain a removal efficiency of 100% for non-volatile contaminants because water penetrates the MD membrane as vapour (Ramlow et al. 2017). The MD method operates at a lower pressure (around 2.5 bar) and produces a high-quality permeate (>98% rejection) with reduced fouling propensity compared to conventional distillation processes (Eykens et al. 2017). Compared with pressure-driven membrane processes (e.g., NF), the minimal vapour pressure differential necessitates less stringent mechanical qualities in the membrane materials (Yan et al. 2019). To date, MD has been the subject of substantial research on brackish water treatment (Ramezanianpour & Sivakumar 2014b; Sivakumar et al. 2014) and desalination (Jia et al. 2021). However, its application in the reclamation of TE remains largely unexplored (Liu et al. 2020; Jeong et al. 2021).

When using a low saline feed, such as TE, organic fouling could become dominant and affect the MD performance. This could be attributed to the significant presence of effluent organic matter (EfOM) (Kumar et al. 2020). Organic fouling is primarily caused by EfOM owing to physical sorption and electrostatic interactions with the hydrophobic MD membrane (Tibi et al. 2021). According to our literature survey, recent studies have been conducted on organic fouling using MD technology (Cho et al. 2018; Charfi et al. 2021). However, organic fouling mechanisms have been predominantly assessed and identified using individual model organic foulants such as humic acid (HA) (Cho et al. 2018; Liu et al. 2018). The findings obtained using model foulants may not be linearly applicable to the fouling caused by real wastewater owing to its complex composition. In this context, the organic fouling of MD membranes in model foulant scenarios versus real wastewater scenarios (i.e., TE) needs to be explored and elucidated. With regards to the EfOM, it has been acknowledged that the main organic foulants found in TE include humic substances (HS) and polysaccharides (Zheng et al. 2019; Charfi et al. 2021). The significance of HS is because of its sorption onto the hydrophobic MD membrane surface (Naidu et al. 2015), whereas polysaccharides, such as alginates, promote a dense fouling layer due to their high compressibility and deformability (Tibi et al. 2021). Previously, Cho et al. (2018) observed the formation of a fouling layer caused by HA deposition, which reduced the direct contact MD (DCMD) flux by 40%. Moreover, Liu et al. (2018) revealed the development of a dense gel layer after exposure to sodium alginate (SA) because of its hydrophobic affinity to the MD membrane, resulting in a rapid flux decline in the early phase of operation. Owing to the interaction of various organics, Qin et al. (2017) revealed that fouling by a mixture of organics was more severe than fouling caused by individual organics. However, to the best of our knowledge, the comparative evaluation of model and real foulants on organic fouling of MD membrane has yet to be systematically investigated.

Although several studies have been conducted on the fouling of MD membranes, they have predominantly focused on the fouling caused by individual foulants (Han et al. 2017; Cho et al. 2018; Couto et al. 2019). Only a few studies have employed real feed (e.g., TE) (Jeong et al. 2021; Ji et al. 2023), probably because the initial research on MD was focused on its desalination performance. Moreover, the majority of fouling experiments in an MD system have been performed using model organics, which should not be expected to mimic the fouling behaviour in the real wastewater scenario (Fortunato et al. 2018). For practical implications, it is necessary to use a real feed for acquiring a better understanding of fouling behaviour among various foulants. Due to the complexity of organics, Yan et al. (2021) called for a comprehensive evaluation and research on the impacts of organics on the MD performance during the treatment of real wastewater. A complex and thick membrane fouling layer may form during the treatment of TE, which is promoted by complex organic interactions. For instance, Kumar et al. (2020) confirmed that the EfOM in TE has 50% reduction on the permeate flux of the MD system. Similarly, Jeong et al. (2021) reported that a 30% flux decline in MD caused by membrane fouling. It is important to note that most MD investigations using real wastewater as a feed have represented membrane fouling as a decline in permeate flux that might not correspond to the complete extent of organic fouling (Naidu et al. 2015; Fortunato et al. 2018; Asif et al. 2021). A valuable resource for understanding organics and their behaviour in membrane fouling could be achieved by employing a more responsive and sensitive analytical approach, such as liquid chromatography–organic carbon detection (LC-OCD). Research on membrane fouling in MD is based on short-term experiments (12–24 h) conducted mostly in the concentration mode (Asif et al. 2021; Ji et al. 2023). To trace the development of organic fouling, it was envisioned to evaluate the long-term performance of DCMD in a continuous-flow system for TE treatment.

Based on the research gaps highlighted above, this study aims to provide an effective understanding of the organic fouling behaviour caused by HS and polysaccharides in a DCMD process, including their rejection. We selected the DCMD configuration owing to its operational simplicity. The DCMD module was operated under three distinct scenarios: (i) individual model organics at different low to high concentrations (20–180 mg/L), (ii) a combination of model organics at concentrations comparable to EfOM in TE (5–20 mg/L each), and (iii) TE collected from a local WWTP. The influence of individual foulants and their combinations on fouling propensity was assessed and compared with that caused by TE. It is essential to note that the study of model organics was conducted in the absence of ions under low-salinity feed conditions. Several characterisation and analytical tools were employed to gain a deep understanding of the correlation between the fouling propensity and flux decline. In this context, HS were given special attention since they could thermally disintegrate into low-molecular-weight (LMW) organics. We believe that the results of this study have practical implications and will serve as a valuable reference for future studies of TE treatment using DCMD.

Chemicals and materials

Two different model foulants of analytical grade were selected (SA and HA) Sigma-Aldrich (St. Louis, MO, USA). These were selected because they are predominantly present in TE to represent polysaccharides and HS. Deionised water was used to prepare stock solutions. Effluent treated biologically was collected from a municipal WWTP (NSW, Australia), and the basic characterisations were carried out. The wastewater had a total organic carbon (TOC) concentration of 40.29 ± 0.1 mg/L, a pH of 6.36, and a conductivity 0.929 mS/cm.

The DCMD experimental setup is depicted in Figure 1, which has the following components: a 3 L glass feed tank, a 5 L permeate tank, a membrane cell, two gear pumps (Micropump Inc., USA), a heating coil (Process Technology, USA), and a temperature regulator (Cole-Parmer, USA). Two blocks of acrylic glass were joined together and were separated by a membrane to form the MD cell. The feed and permeate channels were carved out across the membrane from the inside of the cell. Each channel was 145, 95, and 3 mm in length, width, and depth, respectively. The feed solution was continuously circulated from the feed tank through a membrane cell. A temperature sensor was added before the feed input to the membrane cell to record the feed temperature. The feed solution was kept at a constant temperature (i.e., 50 ± 1.5 °C) using a temperature control device that was connected to a heating coil and a temperature sensor. A second sensor was installed to measure the permeate temperature at the membrane cell outlet. A chiller (SC100-A10, Thermo Scientific, USA) was used in conjunction with a stainless steel heat exchanging coil immersed directly in the permeate tank to maintain the permeate at 20 ± 0.1 °C. An analytical balance (Mettler Toledo, Switzerland) continuously recorded the weight measurements of a glass container into which the overflow from the permeate tank was collected. Water (1 L/min) was pumped through the membrane cell from the feed tank. Meanwhile, 1 L/min of permeate water was pumped into the cell from the permeate tank. Generally, porous membranes such as polyethersulfone (Acarer 2022), polysulfone (Jafar Mazumder et al. 2020), and cellulose acetate (Battirola et al. 2017) are commonly used in filtration applications. In this study, we selected a polytetrafluoroethylene (PTFE)-based hydrophobic MD membrane owing to its higher flux and better chemical stability than other MD membranes under the same operational conditions (Eykens et al. 2017). Anow Microfiltration (China) supplied a flat-sheet PTFE membrane. The manufacturer claimed a pore size of 0.2 μm, thickness of 100 μm, and porosity of 80%. The vapour pressure gradient created by the temperature differential between the feed and the permeate sides of the membrane caused water to move from the feed side to the permeate side, increasing the volume of water in the permeate tank. Under these operating conditions, the pure water flux of DCMD was approximately 10 L/m2 h.
Figure 1

Experimental setup of direct contact MD.

Figure 1

Experimental setup of direct contact MD.

Close modal

Experimental procedures

Impacts of model organics

A total of 13 DCMD experiments were conducted for the model organics, as shown in Table 1. Experiments were conducted with the aim of exploring the impact of model organic fouling over a diverse range of concentrations and combinations. As a model HS, five different concentrations of HA (i.e., 20, 60, 100, 140, and 180 mg/L) were used in the first five runs. Five runs were performed with different concentrations of SA (i.e., 20, 60, 100, 140, and 180 mg/L) as model polysaccharides. Three different combinations of SA and HA were investigated to explore their interactions between HA and SA in the MD system. For these three runs, the total concentration was maintained at 20 mg/L with ratios of HA:SA (25/75, 50/50, and 75/25). The initial volume of the feed was maintained at 3 L for all model organics studies and was conducted in concentration mode with 80% recovery. The feed solution was filtered through a 0.45 μm membrane filter paper to prevent the initial deposition of suspended particles on the MD membrane.

Table 1

Experimental runs for model organics

RunHA (mg/L)SA (mg/L)Duration (min)
20 – 1,040 
60 – 1,090 
100 – 1,140 
140 – 1,190 
180 – 1,290 
– 20 1,140 
– 60 1,270 
– 100 1,400 
– 140 1,490 
10 – 180 1,710 
11 15 1,080 
12 10 10 1,170 
13 15 1,040 
RunHA (mg/L)SA (mg/L)Duration (min)
20 – 1,040 
60 – 1,090 
100 – 1,140 
140 – 1,190 
180 – 1,290 
– 20 1,140 
– 60 1,270 
– 100 1,400 
– 140 1,490 
10 – 180 1,710 
11 15 1,080 
12 10 10 1,170 
13 15 1,040 

Impact of real wastewater (TE)

To elucidate the fouling effect of real wastewater on the membrane, the MD module was run in two distinct modes: (i) concentration mode and (ii) continuous mode. TE was used as the feed in concentration and continuous modes, and the volume of feed tank was 3 L. The MD system was operated at an 80% recovery rate using an identical feed. A concentration mode cycle was conducted to evaluate the impact of the bulk organics in the feed. Ten cycles of the concentration mode were designed to achieve a 60% reduction in permeate flux, which served as the basis for a comparative analysis of the projected performance of a high concentration of model organics. The continuous mode was then studied by maintaining the volume of the feed tank at 3 L, which was controlled by a level sensor connected to a peristaltic pump fed by a TE reservoir tank. This mode was operated for 22 days until the flux decline was 90%. This showed the long-term effect of real wastewater (TE) on MD membrane fouling.

Analytical methods and characterisations

Permeate and feed samples were analysed for aromaticity (UV254) at 254 nm using a UV–visible spectrometer (UV-2600, Shimadzu, Japan). The EfOM was analysed as TOC using a TOC analyser (Shimadzu, Japan). The morphology and composition of the fouling layers on the MD membrane surface were determined by scanning electron microscopy (SEM) combined with energy-dispersive X-ray spectroscopy (EDS) (JCM-600, JEOL, Tokyo, Japan). As the membrane samples were non-conductive, they were coated with platinum (Pt) for SEM-EDS analysis. Using attenuated total reflection-Fourier transform infrared (ATR-FTIR) spectroscopy, the chemical composition of the fouling layer was analysed (IRAffinity-1, Shimadzu, Kyoto, Japan). Using a goniometer (Model250, Rame-Hart, Netcong, NJ) based on the usual sessile drop method, the contact angle was measured. Using LC-OCD, the detailed organic characterisation of the feed, permeate, and the membrane fouling layer was assessed (Model 8, DOC-Labor, Karlsruhe, Germany). The major fractions detected by LC-OCD include biopolymers (BP), building blocks (BB), LMW acids, and neutrals (Naidu et al. 2015; Fortunato et al. 2018; Asif et al. 2021).

Influence of model organic foulants on fouling behaviour

The effects of different HA concentrations on the relative flux decline in the DCMD system were investigated, as shown in Figure 2. Experiments conducted with various concentrations of HA revealed a rapid decline in flux within the first 100 min of DCMD operation (Figure 2(a)). This reduction in the permeate flux correlated well with the concentration of the model foulant HA. The flux decline ranged between 2% (at HA = 20 mg/L) and 25% (at HA = 180 g/L). After the initial flux decline, the permeate flux almost remained constant. Given the nature of low-pressure-driven membrane processes such as MD, it is anticipated that an individual organic feed will not undergo significant physical or chemical interactions owing to its homogeneity. Membrane fouling, on the other hand, is a process wherein various particles engage in physical or chemical interactions with each other and with the membrane surface (Acarer 2023).
Figure 2

Normalised flux vs. operating time at different concentrations of model organics, including (a) HA and (b) SA. Experimental conditions: Feed temperature = 50 ± 1.5 °C; distillate temperature = 20 ± 0.1 °C; circulation rate = 1 L/min; and recovery rate = 80%.

Figure 2

Normalised flux vs. operating time at different concentrations of model organics, including (a) HA and (b) SA. Experimental conditions: Feed temperature = 50 ± 1.5 °C; distillate temperature = 20 ± 0.1 °C; circulation rate = 1 L/min; and recovery rate = 80%.

Close modal

Notably, at HA concentrations of 20, 60, and 100 mg/L, the flux reduction was insignificant and remained below 10% (Figure 2(a)). It is evident from the experiment that when the concentration of the individual HA feed is low, they do not exhibit notable bonding with the membrane surface. This is in accordance with a previous investigation, in which a feed containing 100 mg/L of HA, even in the presence of salts, did not affect the permeate flux, i.e., less than 5% flux decline (Han et al. 2017). Therefore, it can be inferred that high HA concentrations (140 and 180 mg/L) play a significant role in the fouling behaviour of the MD membrane. However, the initial decline in flux for all concentrations of HA was due to the initial deposition of HA on the membrane surface, which was caused by its hydrophobic interactions with the MD membrane surface (Han et al. 2017; Cho et al. 2018). Subsequently, a significant initial flux decline was observed only when the HA concentrations in the feed were 140 and 180 mg/L (Figure 2(a)). Interestingly, this initial flux decline only occurred within the first 100 min of the DCMD operation. This is because HA forms a fouling layer via hydrophobic interactions, which seems to limit further HA deposition on the membrane surface (Couto et al. 2019). A previous study (Cho et al. 2018) showed that with the variation in HA concentration (0–200 mg/L) in feed solutions during short-term experiments, the extent of flux decline was governed by the concentration of HA, which is consistent with the findings of this study.

The behaviour of the flux decline in DCMD by SA (Figure 2(b)) was observed similar to that obtained for different HA concentrations (Figure 2(a)). There was also an initial drop in permeate flux across all SA concentrations. Similarly, the initial decline in flux was comparatively higher (5–60%) for SA concentrations of 140 and 180 mg/L. Moreover, this initial flux decline also occurred within the first 100 min, after which the permeate flux remained almost constant owing to the homogeneity of the feed solution or the hydrophilic nature of the fouling layer. When the flux declines in HA and SA (Figure 2) were compared, the flux decline in SA was greater than that in HA, although both tended to achieve a pseudo-steady state. The presence of carboxylic groups in SA causes it to be deposited on the MD membrane as a gel-like structure (Liu et al. 2018). This trend was observed for all concentrations of the SA feed, indicating that polysaccharides caused far more severe membrane fouling. Our findings concur with the existing literature, indicating that polysaccharides are the primary constituents of fouling in natural water (Katsoufidou et al. 2010). In a study conducted by Kim et al. (2019), it was observed that MD feed containing 50 mg/L of SA did not result in any decline in flux, which aligns with the findings of the present study (Figure 2). In brief, it was observed that an initial higher feed concentration results in a higher permeate flux decline, a phenomenon that appears to occur only during the early stages of DCMD operation. We suggest that the change in membrane hydrophobicity caused by the hydrophilic and negatively charged fouling layer may limit further deposition of model organic foulants after a certain operating time.

Impacts of organic foulant combinations on fouling behaviour

To determine the effects of combinations of organic foulants, we evaluated mixed simulated feed solutions containing HA and SA at different concentrations. The results (Figure 3(a)) showed an almost constant permeate flux with a slight initial flux decline (<5%) when the concentration of either HA or SA was 5 mg/L. When the mass ratio of HA:SA was 1:1, the drop in the permeate flux was almost 10%. Through experimentation, it was ascertained that a combination of organic foulants did not induce significant membrane fouling on the MD membrane. Compared with individual organic foulants (Figure 2), it cannot be conclusively inferred that the combination of organic foulants exacerbated fouling in DCMD under the operating conditions of this study. In contrast to the results of this study, a previous study reported rapid membrane fouling in the case of the HA:SA combination compared to individual foulants (Myat et al. 2014). This contradiction could be linked to the differences in the chemistry or constituents of the synthetic feed. The MD feed in this study did not contain inorganic impurities, such as Ca2+, which can act as a bridge between the organic foulants and the fouling layer (Asif et al. 2021). Hence, severe membrane fouling was not observed in the case of the HA:SA combinations (Figure 3(a)). Specifically, in all cases with HA:SA combinations, after an initial phase of slight reduction in flux lasting approximately 30 min, the permeate flux virtually stabilised. However, the decline was slightly greater than that observed in experiments with individual organics of the same concentration. Furthermore, as previously indicated, this behaviour was due to the formation of a labile layer that was only weakly linked to the membrane by the interaction between SA and HA. The HA:SA (1:1 mass%) combination as a feed exhibited the most rapid flux decline; the flux dropped to 90% within the first 50 min. The HA:SA combinations appeared to follow the same pattern. For instance, in the event of a mixture containing an equal proportion of humic compounds and polysaccharides, fouling was considerably more severe. This phenomenon was previously illustrated by Katsoufidou et al. (2010), who demonstrated that HA molecules tend to become entrapped inside the deposited mass of SA, indicating that SA acts as a binder. Nevertheless, it is intriguing that this behaviour was observed in the presence of both organics (HA and SA). Consequently, it appears that equal concentrations of HA and SA can form a denser cake layer on the membrane surface.
Figure 3

Normalised flux vs. operating time during treatment of (a) simulated wastewater containing different combinations of model foulants and (b) long-term performance of DCMD for the treatment of real TE wastewater in different operating modes. Unless otherwise stated in the graphs, the experimental conditions were as follows: feed temperature = 50 ± 1.5 °C; permeate temperature = 20 ± 0.1 °C; circulation rate = 1 L/min; and recovery rate = 80%.

Figure 3

Normalised flux vs. operating time during treatment of (a) simulated wastewater containing different combinations of model foulants and (b) long-term performance of DCMD for the treatment of real TE wastewater in different operating modes. Unless otherwise stated in the graphs, the experimental conditions were as follows: feed temperature = 50 ± 1.5 °C; permeate temperature = 20 ± 0.1 °C; circulation rate = 1 L/min; and recovery rate = 80%.

Close modal

Impacts of real wastewater (TE) on fouling behaviour

To investigate the applicability of the fouling behaviour assessed using model organic foulants and actual wastewater, the DCMD was operated with real feed (i.e., TE) in the concentration and continuous modes. The permeate flux results for the real feed (Figure 3(b)) showed a significantly different trend from that of the model organics. The gradual decrease in the permeate flux observed in the TE samples was not present in the model organics, even at high concentrations. This disparity can be attributed to the presence of complex organics in the TE, which led to the continuous deposition of EfOM. Moreover, the deposition of foulants is facilitated by inorganics even when the ionic strength of the feed is very low. The findings of this study regarding the DCMD flux behaviour for wastewater treatment are consistent with those reported in several previous studies (El-Abbassi et al. 2013; Lin et al. 2015).

The EfOM in TE comprises complex organics (Manikandan et al. 2022), which results in a significant decline in the permeate flux. Moreover, compared to model organics, TE also contains ions (such as Mg2+ and Ca2+) that interact with the negatively charged organics adsorbed on the membrane surface and may neutralise the ions (Yan et al. 2019). It is well known that divalent scale-forming cations act as a bridge between organic foulants, providing active sites for organic foulant attachment or disposition. This resulted in the formation of a dense membrane fouling layer. In addition, complex organic foulants can form a cation-stabilised gel, termed a Begg-box-shaped gel (Ji et al. 2023). Subsequently, the scaling agents might be further attached via electrostatic interactions to an already fouled layer, turning the membrane surface even less negatively charged. This continuous deposition significantly lost the hydrophobicity of the membrane surface. Despite the severe and steady build-up of the fouling layer, the DCMD was able to remove more than 98% of organics, as evident from the LC-OCD analysis.

The concentrate mode experiments revealed a steeper flux decline during all 10 runs over a 9-day period (Figure 3(b)). In contrast, the long-term study performed in continuous mode demonstrated a gradual decline in flux for the first 10 days. A slight increase in flux decline was observed between days 10 and 18. Furthermore, an abrupt flux decline was observed between days 18 and 21. However, the flux remained stable between days 21 and 25. As reported by others, this trend is caused by the breakdown of high-molecular-weight organics into low-molecular-weight organics between days 18 and 21. The deposition of organics with LMW was evident based on the LC-OCD analysis, as explained in subsequent sections. It is possible that small particles might penetrate or adhere to the membrane surface, which is likely the cause of the substantial reduction in permeate flux between days 18 and 21 (Liu et al. 2020). In summary, it can be deduced that the fouling behaviour during the treatment of real wastewater differs significantly from that observed in the case of synthetic wastewater containing model foulants.

Membrane autopsy

Morphology and composition of fouling layer

The morphological features and composition of the fouling layer formed during the different experiments were analysed using SEM-EDS (Figure 4). Carbon (C) and oxygen (O) peaks characterise organic fouling deposits, which are based on prior membrane fouling studies that analysed organic fouling through C and O element peaks (Naidu et al. 2014). A peak representing Pt was observed, which was used for the coating of the membrane samples. The surface of the virgin PTFE membrane exhibited a typical microporous nature (Figure 4(a)), which was dominated by C (10.3%) and F (71.6%), whereas the presence of O (0.5%) could be an impurity or noise. Depending on the feed composition, the fouled MD membrane samples revealed that the membrane was covered with organic materials. This reduces the active surface area of the membrane for water vapour transfer from the feed to the permeate (Asif et al. 2021), thereby decreasing the permeate flux. As measured by SEM-EDS, the membrane fouled by SA had a C peak with higher intensity than that observed in the case of the virgin MD membrane and the membrane fouled by HA (Figure 4). Although the SA foulant layer appeared invisible to the naked eye, a foulant layer nonetheless developed on the membrane surface, which is consistent with a previous study (Liu et al. 2018). Sample IDs, including SA-180, SA-140, SA-100, SA-60, and SA-20, were 14.3, 14.1, 13.9, 13.8, and 11.4% carbon, respectively, which further verifies that organic foulants accumulated on the surface of the MD membrane. The first deposition of organics was dependent on the initial concentration, indicating that after the initial attachment, individual foulants did not adhere to the surface because of hydrophobic interactions in the absence of bridging impurities. Similarly, the membrane samples after the experiment ID of HA-180, HA-140, HA-100, HA-60, and HA-20 had 13.9, 11.5, 11.4, 11.2, and 11% carbon, respectively. A comparison of the membranes before and after the DCMD experiment of HA demonstrated the emergence of darker zones owing to the formation of brownish dense colloids, which are visible to the naked eye. The combination of organics resulted in HA and SA deposition, although a comparable effect was observed with respect to carbon content (%). Carbon contents of 12.8, 12.5, and 11.9% were found in the case of HA:SA ratios of 10:10, 5:15, and 15:5, respectively. It should be noted that the percent carbon content in the MD membrane after the experiments using model organic foulants remained closer to that of the virgin membrane. This was because the model organic foulants were deposited on the MD membrane surface unevenly and did not form a dense fouling layer. Therefore, after reduction during the initial phase, the permeate flux of the DCMD remained stable during all the experiments using individual or combinations of model organic foulants.
Figure 4

SEM-EDS analysis of MD membranes before and after fouling experiments. (a) Virgin membrane, (b) MD membrane fouled by SA (180 mg/L), (c) MD membrane fouled by SA (140 mg/L), (d) MD membrane fouled by SA (100 mg/L), (e) MD membrane fouled by SA (60 mg/L), (f) MD membrane fouled by SA (20 mg/L), (g) MD membrane fouled by HA (180 mg/L), (h) MD membrane fouled by HA (140 mg/L), (i) MD membrane fouled by HA(100 mg/L) (j) MD membrane fouled by HA (60 mg/L), (k) MD membrane fouled by HA (20 mg/L), (l) MD membrane fouled by SA (15 mg/L)/HA (5 mg/L), (m) MD membrane fouled by SA (10 mg/L)/HA (10 mg/L), (n) SA (5 mg/L)/HA (15 mg/L), and (o) MD membrane fouled by TE in continuous mode.

Figure 4

SEM-EDS analysis of MD membranes before and after fouling experiments. (a) Virgin membrane, (b) MD membrane fouled by SA (180 mg/L), (c) MD membrane fouled by SA (140 mg/L), (d) MD membrane fouled by SA (100 mg/L), (e) MD membrane fouled by SA (60 mg/L), (f) MD membrane fouled by SA (20 mg/L), (g) MD membrane fouled by HA (180 mg/L), (h) MD membrane fouled by HA (140 mg/L), (i) MD membrane fouled by HA(100 mg/L) (j) MD membrane fouled by HA (60 mg/L), (k) MD membrane fouled by HA (20 mg/L), (l) MD membrane fouled by SA (15 mg/L)/HA (5 mg/L), (m) MD membrane fouled by SA (10 mg/L)/HA (10 mg/L), (n) SA (5 mg/L)/HA (15 mg/L), and (o) MD membrane fouled by TE in continuous mode.

Close modal

When the DCMD was run with real wastewater (i.e., TE), a dense fouling layer was formed. The SEM images revealed that the organics significantly affected the morphology of the fouling layer. EDS was used to detect the presence of many elements (i.e., C, O, P, Cl, Fe, S, Na, Ca, Mg, K, and Si) in the fouling layer. With TE as a feed, the fouling layer was mostly composed of O (54%) and Ca (11.3%), whereas the cumulative concentration (% by weight) of other components, including P, Cl, Na, S, and Mg, was less than 10%. This proves that inorganic foulants, predominantly Ca2+, act as a bridge between organic foulants and hydrophobic membrane, resulting in severe and rapid organic fouling in MD (Liu et al. 2018). The SEM images showed that the deposits were extremely dense on the membrane surface. The insignificant presence of divalent ions (e.g., Ca2+ and Mg2+) can result in the formation of metal–natural organic matter complexes, which may contribute to highly compacted fouling deposits and a decrease in membrane performance (Silva et al. 2018). Based on the results of the EDS analysis of the fouled membrane, calcium (Ca) was present, which promoted the formation of calcium precipitates. The rapid conversion of HCO3 to CO32− in the feed solution under heating conditions could expedite calcium deposition on the MD membrane (Tijing et al. 2015). In conclusion, HA and SA compounds individually resulted in initial (uneven) deposition governed by their initial concentrations because of the interaction between the hydrophobic membrane and the organics. Even a combination of model organics did not result in considerable fouling development. Considerable fouling development was observed in MD when TE was used as the feed. Considering the distinct organic fouling patterns observed on hydrophobic membranes, this study conducted additional research with a feed solution and characterised organic foulants. This is especially pertinent because most real feeds contain complex organic compounds.

Effect of fouling layer on membrane hydrophobicity

The deposition of organic and inorganic foulants is anticipated to influence the membrane surface hydrophobicity. Therefore, the water contact angles of the virgin and fouled MD membranes were measured. There was an obvious loss in membrane hydrophobicity because of the deposition of membrane foulants during the different feed composition experiments (Figure 5). The virgin PTFE membrane was hydrophobic and had a high contact angle of 135°, whereas the membranes after the DCMD operation using model foulants exhibited contact angles in the range of 60°–90° (Figure 5). For instance, the MD membranes fouled by HA and SA showed loss of hydrophobicity, with contact angles of 70.6° and 60.2°, respectively. Notably, HA and SA deposition increased the negative charge on the membrane surface owing to their hydrophilic nature (Liu et al. 2018; Couto et al. 2019). A membrane hydrophobicity loss was also observed in the combined organics experiment (i.e., SA + HA), with a contact angle of 65.5°. In the interaction between SA and HA molecules, some HA molecules tend to be trapped in the deposited mass of SA on the membrane surface. This causes a formation of a cake layer on the membrane surface where SA acts as a binder (Katsoufidou et al. 2010). Therefore, the experiments involving equal proportions of HA and SA resulted in more hydrophilic compared to the other combinations. In contrast, the contact angles observed for the membranes used for the TE feed were 27.2° and 25.5° following operation in the concentration and continuous modes, respectively. The significantly high hydrophilicity of the membrane may be attributed to the deposition of complex organics and inorganics on the membrane surface, generating a dense fouling layer. Furthermore, the decomposed HS in the TE (discussed in Section 3.4.4) could interact with the hydrophobic membrane, leading to a significant loss of hydrophobicity. Throughout all the studies, the feed and permeate analyses revealed 98% rejection of organics, indicating that pore wetting did not occur. However, this loss in hydrophobicity might increase the probability of membrane wetting. In addition, the impact of pore wetting on the MD membrane is more noticeable at higher feed temperatures (Ramezanianpour & Sivakumar 2014b; Sivakumar et al. 2014). Consistent with the findings of this experiment, it was reported that fouling of the MD membrane during wastewater treatment lost the membrane hydrophobicity (with a contact angle of 60°); besides, no pore wetting was observed (Song et al. 2018). In another study, Lee et al. (2021) showed that the hydrophobicity was lost and the contact angle reduced to 20° first, followed by another reduction to 0°.
Figure 5

Effect of membrane fouling layer on hydrophobicity of the MD membrane after fouling experiment conducted using model foulants and real wastewater.

Figure 5

Effect of membrane fouling layer on hydrophobicity of the MD membrane after fouling experiment conducted using model foulants and real wastewater.

Close modal

Effect of fouling layer on membrane surface functionalities

ATR-FTIR analysis was performed on virgin and fouled MD membranes to analyse the functional groups of the membrane foulants (Figure 6). Changes in the chemical surface groups were studied by recording and evaluating the FTIR spectra of each membrane. As shown in Figure 6, the stretching vibration of the CF2 group displayed peaks at 1,202 and 1,147 cm−1 for the virgin PTFE membrane, which is also evident in other studies (Liu et al. 2018; Ji et al. 2023). In general, the FTIR absorbance of the typical absorption peaks decreased as the organic content increased. In the case of the fouling layer formed by the model organics, several additional peaks were observed at 3,691, 2,927, 1,543, 1,040, and 915 cm−1 (Figure 6(a)–6(c)). The peaks between 2,900 and 3,600 cm−1 correspond to O–H and C–H stretching vibrations, respectively (Guo et al. 2020). On the other hand, the C = O and C–H bonds are represented by peaks in the 910–1,050 cm−1 region. Importantly, the membranes fouled by the TE feed exhibited a very strong peak at 1,396 cm−1 (Figure 6(d)), which was not observed in the case of model foulants (Figure 6(a)–6(c)). The peak at 1,396 cm−1 corresponds to the stretching of Amide III or COO side chain bonds in proteins (1,200–1,400 cm−1) (Nara et al. 2013). A similar peak (1,402 cm−1) was reported for a polyvinylidene fluoride membrane contaminated with algal membrane bioreactor effluent (Sun et al. 2021). The peak at 870 cm−1 corresponds to C–H bending bonds in polysaccharides (Song et al. 2018). A comparison of the functional groups in the fouling layer formed on the MD membrane indicated that the membrane fouling caused by real wastewater completely covered the MD membrane surface, as indicated by the disappearance of its characteristic peaks. In addition, proteins and polysaccharides in real wastewater feed seem to be the dominant components of the fouling layer as compared to HS. Our findings clearly show that the fouling behaviour observed or investigated using model organics could not be linearly applied to the fouling caused by real wastewater. Therefore, future studies must adopt the practice of utilising real wastewater as feed to understand and delineate the fouling behaviour and mechanisms.
Figure 6

Effect of membrane fouling layer on MD membrane functionalities using model foulants such as HA (a), SA (b), and HA/SA (c) as well as real wastewater (d).

Figure 6

Effect of membrane fouling layer on MD membrane functionalities using model foulants such as HA (a), SA (b), and HA/SA (c) as well as real wastewater (d).

Close modal

Characterisation of organic foulants by LC-OCD

The LC-OCD technique was employed to characterise the EfOM in the TE, membrane foulants, and DCMD permeate based on the molecular weight distributions, as shown in Figure 7. The molecular weight of organic matter (EfOM) affects the retention time (RT) for elution from the LC-OCD size exclusion column. In the LC-OCD analysis with the column settings used in this study, an RT of 30–40 min corresponded to BP with molecular weights greater than 20 kDa, whereas HS and BB RT were 60 and 70 min with molecular weights in the range of 1–20 kDa and 350–500 Da, respectively. Moreover, LMW acids and neutrals are characterised by a molecular size distribution of less than 350 Da eluted with an RT with a range of 70–90 and 100–120 min, respectively. The LC-OCD chromatogram of the TE feed revealed three main organic constituents: HS, LMW neutrals, and BB (Figure 7). The analysis showed that the distribution of EfOM in the TE was HS (27.3%), LMW neutrals (12.5%), and BB (9.5%). A small fraction of BP (3%) was also observed. The largest EfOM fraction in the TE was HS (i.e., 21.4 mg/L), which was also reported by Siebdrath et al. (2021).
Figure 7

LC-OCD chromatograms of TE feed, permeate and membrane fouling layer.

Figure 7

LC-OCD chromatograms of TE feed, permeate and membrane fouling layer.

Close modal

The chromatogram obtained from the LC-OCD analysis of the DCMD permeate demonstrated notable rejection of EfOM. The permeate TOC concentration was found to be 1.23 mg/L, indicating a retention of 98% organics by the MD membrane. The MD process is supposed to only allow volatiles to pass; however, as shown in Figure 7, the prevalent organic compound present in the permeate was HS, constituting 36.8% of the total organics, with a concentration of 0.4 mg/L. This is due to the reason that HS can pass through the membrane to permeate via adsorption–desorption inside the unwetted membrane pores (Li et al. 2023). The overall results of LC-OCD in the DCMD permeate showed LMW neutrals (0.21 mg/L, 17%), BB (0.12 mg/L, 10%), and BP (0.09 mg/L, 7.4%). This shows that there was a high rejection of organics by the PTFE membrane and no pore wetting occurred.

To further identify the organics which lead to fouling, LC-OCD analysis was performed on the membrane foulant layer. The LC-OCD chromatogram of the membrane foulants shown in Figure 7 shows a high peak at an RT of 110 min, which indicates a high deposition of LMW neutrals and BB on the membrane surface. The presence of LMW neutrals and BB components in the total organic deposition was reported at 29.8 and 25%, respectively. Surprisingly, LMW organics were not considerably present in the TE feed, this phenomenon has previously been reported as high feed temperatures thermally disaggregate HS to LMW compounds (Fortunato et al. 2018). In membrane filtration, LMW neutrals have been recognised as one of the most prominent organic foulants (Li et al. 2016). The hydrophobicity of the MD membrane promoted the adsorption of organic materials on its surface. According to Zheng et al. (2019), proteins tend to accumulate on hydrophobic membranes (Naidu et al. 2014).

The TE feed contains complex organics with both hydrophilic and hydrophobic molecules (Naidu et al. 2014). The LC-OCD results showed that there were 47.5% (37.31 mg/L) hydrophobic and 52.5% (41.20 mg/L) hydrophilic compounds in the TE feed. The interaction between the hydrophobic and hydrophilic organic compounds forms a mechanism for the layer-by-layer effect. Initially, hydrophobic molecules are adsorbed onto the membrane surface, followed by the accumulation of hydrophilic organics, resulting in a dense layer on the membrane surface, leading to flux decline. Naidu et al. (2017) demonstrated that LMW organics possess hydrophilic propensities causing a substantial reduction in membrane hydrophobicity, causing a significant deposition of hydrophilic LMW organics. Moreover, Zheng et al. (2021) demonstrated the hydrophobic–hydrophilic interaction as ‘brick-layering’ process, where hydrophobic foulants acts as concrete and hydrophilic foulants act as bricks. Thus, it can be concluded that model organics (such as HA and SA) do not accurately replicate real humic compounds and polysaccharides.

This study was carried out to better understand the fouling behaviour of organic foulants in a direct contact MD (DCMD) system. To assess the impact of organic fouling on the MD process, a comparison between the model and real organics was performed. The findings presented here indicate that, in the DCMD system, model organics behave differently from real organics. Even a high concentration of individual model organics only caused an initial flux decline, after which the flux became stable. Moreover, there was no considerable decline in the flux with the combination of model organics. However, the real organics (EfOM) in the TE contained complex organics, including hydrophobic and hydrophilic organics. Organic fouling development using TE as the feed is a phenomenon of hydrophobic–hydrophilic interaction, which is initiated by the deposition of LMW hydrophobic organics, followed by interaction with hydrophilic organics. Thus, model organics may not be directly used in future research. Moreover, the DCMD system was also proven to be an effective way to reclaim water from the TE, as evident from the permeate quality (>98% organic rejection). The results of this study will not only aid future investigations into wastewater reclamation by DCMD systems but will also contribute to the development of high retention membrane technology for wastewater reclamation. Further investigation of the impacts of biofouling and pretreatment options using real wastewater is also recommended.

This research was carried out with the support of the Australian Government Research Training Program Scholarship to RH through the University of Wollongong (UOW). The authors acknowledge the use of facilities available within the School of Civil, Mining, Environmental and Architectural Engineering of UOW. We acknowledge access to FTIR at the School of Chemistry and Molecular Bioscience of UOW, Australia.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Abbas
T. K.
,
Rashid
K. T.
,
Al-Saady
S.
,
Alsarayrehd
A. A.
,
Figoli
A.
&
AlSalhy
Q. F.
2023
Decontamination of aqueous nuclear waste via pressure-driven membrane application – A short review
.
Engineering and Technology Journal
41
(
9
),
1152
1174
.
Battirola
L. C.
,
Andrade
P. F.
,
Marson
G. V.
,
Hubinger
M. D.
&
Gonçalves
M. d. C.
2017
Cellulose acetate/cellulose nanofiber membranes for whey and fruit juice microfiltration
.
Cellulose
24
(
12
),
5593
5604
.
Cho
D.-W.
,
Kwon
G.
,
Han
J.
&
Song
H.
2018
Influence of humic acid on the long-term performance of direct contact membrane distillation
.
Energy & Environment
30
(
1
),
109
120
.
Couto
C. F.
,
Amaral
M. C. S.
,
Lange
L. C.
&
Santos
L. V. d. S.
2019
Effect of humic acid concentration on pharmaceutically active compounds (PhACs) rejection by direct contact membrane distillation (DCMD)
.
Separation and Purification Technology
212
,
920
928
.
El-Abbassi
A.
,
Hafidi
A.
,
Khayet
M.
&
García-Payo
M. C.
2013
Integrated direct contact membrane distillation for olive mill wastewater treatment
.
Desalination
323
,
31
38
.
Eykens
L.
,
De Sitter
K.
,
Dotremont
C.
,
Pinoy
L.
&
Van der Bruggen
B.
2017
Membrane synthesis for membrane distillation: A review
.
Separation and Purification Technology
182
,
36
51
.
Fortunato
L.
,
Jang
Y.
,
Lee
J.-G.
,
Jeong
S.
,
Lee
S.
,
Leiknes
T.
&
Ghaffour
N.
2018
Fouling development in direct contact membrane distillation: Non-invasive monitoring and destructive analysis
.
Water Research
132
,
34
41
.
Han
L.
,
Xiao
T.
,
Tan
Y. Z.
,
Fane
A. G.
&
Chew
J. W.
2017
Contaminant rejection in the presence of humic acid by membrane distillation for surface water treatment
.
Journal of Membrane Science
541
,
291
299
.
Jafar Mazumder
M. A.
,
Raja
P. H.
,
Isloor
A. M.
,
Usman
M.
,
Chowdhury
S. H.
,
Ali
S. A.
,
Inamuddin
&
Al-Ahmed
A.
2020
Assessment of sulfonated homo and co-polyimides incorporated polysulfone ultrafiltration blend membranes for effective removal of heavy metals and proteins
.
Scientific Reports
10
(
1
),
7049
.
Jeong
S.
,
Song
K. G.
,
Kim
J.
,
Shin
J.
,
Maeng
S. K.
&
Park
J.
2021
Feasibility of membrane distillation process for potable water reuse: A barrier for dissolved organic matters and pharmaceuticals
.
Journal of Hazardous Materials
409
,
124499
.
Jia
X.
,
Li
K.
,
Wang
B.
,
Zhao
Z.
,
Hou
D.
&
Wang
J.
2021
Membrane cleaning in membrane distillation of reverse osmosis concentrate generated in landfill leachate treatment
.
Water Science and Technology
85
(
1
),
244
256
.
Katsoufidou
K. S.
,
Sioutopoulos
D. C.
,
Yiantsios
S. G.
&
Karabelas
A. J.
2010
UF membrane fouling by mixtures of humic acids and sodium alginate: Fouling mechanisms and reversibility
.
Desalination
264
(
3
),
220
227
.
Kim
B.
,
Cho
H.
,
Choi
J.
,
Sohn
J.
&
Lee
S.
2019
Fouling-induced wetting of membranes in membrane distillation (MD) systems
.
Desalination and Water Treatment
157
,
360
371
.
Kumar
R. V.
,
Barbosa
M. O.
,
Ribeiro
A. R.
,
Morales-Torres
S.
,
Pereira
M. F. R.
&
Silva
A. M. T.
2020
Advanced oxidation technologies combined with direct contact membrane distillation for treatment of secondary municipal wastewater
.
Process Safety and Environmental Protection
140
,
111
123
.
Manikandan
S.
,
Subbaiya
R.
,
Saravanan
M.
,
Ponraj
M.
,
Selvam
M.
&
Pugazhendhi
A.
2022
A critical review of advanced nanotechnology and hybrid membrane based water recycling, reuse, and wastewater treatment processes
.
Chemosphere
289
,
132867
.
Naidu
G.
,
Jeong
S.
,
Kim
S.-J.
,
Kim
I. S.
&
Vigneswaran
S.
2014
Organic fouling behavior in direct contact membrane distillation
.
Desalination
347
,
230
239
.
Naidu
G.
,
Jeong
S.
&
Vigneswaran
S.
2015
Interaction of humic substances on fouling in membrane distillation for seawater desalination
.
Chemical Engineering Journal
262
,
946
957
.
Naidu
G.
,
Jeong
S.
,
Choi
Y.
&
Vigneswaran
S.
2017
Membrane distillation for wastewater reverse osmosis concentrate treatment with water reuse potential
.
Journal of Membrane Science
524
,
565
575
.
Nara
M.
,
Morii
H.
&
Tanokura
M.
2013
Coordination to divalent cations by calcium-binding proteins studied by FTIR spectroscopy
.
Biochimica et Biophysica Acta (BBA) – Biomembranes
1828
(
10
),
2319
2327
.
Qin
W.
,
Zhang
J.
,
Xie
Z.
,
Ng
D.
,
Ye
Y.
,
Gray
S. R.
&
Xie
M.
2017
Synergistic effect of combined colloidal and organic fouling in membrane distillation: Measurements and mechanisms
.
Environmental Science: Water Research & Technology
3
(
1
),
119
127
.
Ramezanianpour
M.
&
Sivakumar
M.
2014a
An analytical flux decline model for membrane distillation
.
Desalination
345
,
1
12
.
Ramezanianpour
M.
&
Sivakumar
M.
2014b
Energy evaluation and treatment efficiency of vacuum membrane distillation for brackish water desalination
.
Journal of Water Reuse and Desalination
5
(
2
),
119
131
.
Ramlow
H.
,
Machado
R. A. F.
&
Marangoni
C.
2017
Direct contact membrane distillation for textile wastewater treatment: A state of the art review
.
Water Science and Technology
76
(
10
),
2565
2579
.
Siebdrath
N.
,
Skibinski
B.
,
Abraham
S.
,
Bernstein
R.
,
Berger
R.
,
Stein
N.
,
Kaufman
Y.
,
Lerch
A.
&
Herzberg
M.
2021
Impact of pretreatment on RO membrane organic fouling: Composition and adhesion of tertiary wastewater effluent organic matter
.
Environmental Science: Water Research & Technology
7
(
4
),
775
788
.
Silva
T. L. S.
,
Morales-Torres
S.
,
Esteves
C. M. P.
,
Ribeiro
A. R.
,
Nunes
O. C.
,
Figueiredo
J. L.
&
Silva
A. M. T.
2018
Desalination and removal of organic micropollutants and microorganisms by membrane distillation
.
Desalination
437
,
121
132
.
Sivakumar
M.
,
Ramezanianpour
M.
&
O'Halloran
G.
2014
Brackish water treatment for reuse using vacuum membrane distillation process
.
Water Supply
15
(
2
),
362
369
.
Song
X.
,
Luo
W.
,
McDonald
J.
,
Khan
S. J.
,
Hai
F. I.
,
Price
W. E.
&
Nghiem
L. D.
2018
An anaerobic membrane bioreactor–membrane distillation hybrid system for energy recovery and water reuse: Removal performance of organic carbon, nutrients, and trace organic contaminants
.
Science of The Total Environment
628–629
,
358
365
.
Tijing
L. D.
,
Woo
Y. C.
,
Choi
J.-S.
,
Lee
S.
,
Kim
S.-H.
&
Shon
H. K.
2015
Fouling and its control in membrane distillation – A review
.
Journal of Membrane Science
475
,
215
244
.
Yan
Z.
,
Yang
H.
,
Qu
F.
,
Zhang
H.
,
Rong
H.
,
Yu
H.
,
Liang
H.
,
Ding
A.
,
Li
G.
&
Van der Bruggen
B.
2019
Application of membrane distillation to anaerobic digestion effluent treatment: Identifying culprits of membrane fouling and scaling
.
Science of The Total Environment
688
,
880
889
.
Yan
Z.
,
Qu
F.
,
Liang
H.
,
Yu
H.
,
Pang
H.
,
Rong
H.
,
Fan
G.
&
Van der Bruggen
B.
2021
Effect of biopolymers and humic substances on gypsum scaling and membrane wetting during membrane distillation
.
Journal of Membrane Science
617
,
118638
.
Zheng
L.
,
Price
W. E.
&
Nghiem
L. D.
2019
Effects of fouling on separation performance by forward osmosis: The role of specific organic foulants
.
Environmental Science and Pollution Research
26
(
33
),
33758
33769
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).