You are on page 1of 17

Chemical Engineering Journal 235 (2014) 8399

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Review

A review of the production and applications of waste-derived volatile fatty acids


Wee Shen Lee, Adeline Seak May Chua , Hak Koon Yeoh, Gek Cheng Ngoh
Department of Chemical Engineering, Faculty of Engineering, University of Malaya, Lembah Pantai, 50603 Kuala Lumpur, Malaysia

h i g h l i g h t s
 Different types of wastes used for VFA production are reviewed.  Important factors inuencing VFA production performances are detailed.  Various applications of waste-derived VFA are examined.  Effects of VFA chain-length on the performances of the applications are elucidated.  Future research needs of microbial VFA production are proposed.

a r t i c l e

i n f o

a b s t r a c t
Low cost production of volatile fatty acids (VFA) from waste by acidogenic fermentation has drawn extensive research interests as VFA is a critical substrate for microorganisms involved in the production of biodegradable plastics and bioenergy, as well as those in biological nutrient removal processes. This article reviews the various wastes amenable to VFA production, the pertinent factors inuencing the VFA production, and the various applications of the resulting VFA. In addition to the usual need for reasonably high concentration, a key feature for many applications is the distribution of the chain length of the VFA. Means to regulate these performance indicators are surveyed and discussed in detail. 2013 Elsevier B.V. All rights reserved.

Article history: Received 11 March 2013 Received in revised form 15 August 2013 Accepted 1 September 2013 Available online 8 September 2013 Keywords: Volatile fatty acid Acidogenic fermentation Polyhydroxyalkanoate Microbial fuel cell Hydrogen Biological nutrient removal

Contents 1. 2. 3. 4. 5. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Types of wastes for VFA production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Pretreatment of solid wastes for VFA production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Anaerobic technologies for VFA production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Factors influencing VFA production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1. pH. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3. Retention time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1. Hydraulic retention time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 86 86 88 89 89 90 90 91

Abbreviations: 3HB, 3-hydroxybutyrate; 3HV, 3-hydroxyvalerate; BNR, biological nutrient removal; CE, coulombic efciency; CSTR, continuous stirred-tank reactor; COD, chemical oxygen demand; DOC, dissolved organic carbon; FVS, lterable volatile solids; HRT, hydraulic retention time; MFC, microbial fuel cell; MSW, municipal solid waste; OLR, organic loading rate; OFMSW, organic fraction of municipal solid waste; P(3HB), poly(3-hydroxybutyrate); PD, power density; PHA, polyhydroxyalkanoates; PS, primary sludge; sCOD, soluble chemical oxygen demand; SDS, sodium dodecyl sulfate; SDBS, sodium dodecyl benzene sulfonate; SRT, solids retention time; TCOD, total chemical oxygen demand; TOC, total organic carbon; TS, total solids; TVS, total volatile solids; TSS, total suspended solids; UASB, upow anaerobic sludge blanket; WAS, waste activated sludge; VDS, volatile dissolved solids; VFA, volatile fatty acids; VS, volatile solids; VSS, volatile suspended solids. Corresponding author. Tel.: +60 3 79675291; fax: +60 3 79675319. E-mail address: adeline@um.edu.my (A.S.M. Chua). 1385-8947/$ - see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.cej.2013.09.002

84

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

6.

7.

5.3.2. Solids retention time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4. Organic loading rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.5. Additives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Applications of waste-derived VFA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1. Polyhydroxyalkanoates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2. Bioenergy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.1. Electricity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.2. Biogas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.3. Hydrogen. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2.4. Lipids for biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.3. Biological nutrient removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

91 91 92 92 92 93 93 94 94 95 95 96 96 96

1. Introduction The rapid growth of human population and the global economy has led to massive waste generation. Proper waste management is crucial to minimize further environmental degradation and to foster the transition to a sustainable society. The conventional waste management approach is treatment-oriented, which mainly focuses on meeting environmental regulations. This approach neglects the potential of diverting the waste as feedstock for the production of value-added chemicals, as such also reducing the quantities of waste. Therefore, a more enlightened waste management approach is resource recovery, which allows simultaneous minimization of waste and generation of value-added products. Among the various possibilities of the latter, the focus of this review is the production of volatile fatty acids (VFA) from various organic wastes. VFA are short-chain fatty acids consisting of six or fewer carbon atoms which can be distilled at atmospheric pressure [1]. These acids have a wide range of applications such as in the production of bioplastics [2], bioenergy [3,4] and the biological removal of nutrient from wastewater [5]. At present, commercial production of VFA is mostly accomplished by chemical routes [6]. However, the use of non-renewable petrochemicals as the raw materials and the increasing price of oil have renewed the interest in biological routes of VFA production [7]. In biological VFA production, pure sugars such as glucose and sucrose have been commonly employed as the main carbon source [8,9], which raises the ethical concern on the use of food to produce chemicals. This issue can be resolved by utilizing organic-rich wastes such as sludge generated from wastewater treatment plant, food waste, organic fraction

of municipal solid waste and industrial wastewater for VFA production. Such transformation of waste into VFA also provides an alternative route to reduce the ever increasing amount of waste generated. In general, the production of VFA from waste is an anaerobic process involving hydrolysis and acidogenesis (the latter is also known as acidogenic fermentation [10] or dark fermentation [11]), as illustrated in Fig. 1. In hydrolysis, complex organic polymers in waste are broken down into simpler organic monomers by the enzymes excreted from the hydrolytic microorganisms. Subsequently, acidogens ferment these monomers into mainly

Types of organic-rich wastes

Pretreatment of waste

Anaerobic technologies for VFA production Operating conditions for VFA production - pH - Temperature - Hydraulic retention time - Solids retention time - Organic loading rate - Additives

Complex polymers in waste (Polysaccharides, proteins and lipids) Hydrolysis Simpler monomers (Monosaccharides, amino acids and long chain fatty acids) Acidogenesis/ acidogenic fermentation/ dark fermentation Volatile fatty acids (e.g. acetic, propionic, butyric acids)
Fig. 1. Production of volatile fatty acids from waste (adapted from [12,152]).

Volatile fatty acids (VFA)

Applications - Polyhydroxyalkanoates - Electricity - Biogas - Hydrogen - Lipids for biodiesel - Biological nutrient removal
Fig. 2. Outline of the review article: production of waste-derived volatile fatty acids and their applications.

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 Table 1 Various wastes used for the production of VFA. Type of wastes Solid waste Waste activated sludge Organic content (mg COD/L) 5470a 18,657 18,657 14,878 14,890 22,838 20,631 343c Not available 91,900 146,100 166,180 347,000g 196,700g 150,600g Cattail Liquid waste Palm oil mill efuent Not available 88,000 30,600 Olive oil mill efuent 70,400 37,000 11,110 11,110 Paper mill efuent 7740 26,300 8750h 4590 4420 4000 12,000 Gelatin-rich proteinaceous wastewater Pharmaceutical wastewater Mixture of two types of wastes Primary sludge + waste activated sludge Primary sludge + starch-rich wastewater 4000 40,00060,000 Reactor type and operating conditions VFA production performance 2561 mg TOC/L 298 mg COD/g VSS 368 mg COD/g VSS 339 mg COD/L 191 mg COD/L 85 mg COD/g VSS 60 mg COD/g VSS/d 31 mg/g VSS/d 25,000 mg/L 8950 mg COD/L 5610 mg COD/L 36,000 mg/L 40 mg/g VS fed 23,110 mg/L 19,581 mg/L 430 mg/g VS 15,300 mg/L 4100 mg/L/d (as acetic acid) 15,600 mg COD/L 10,700 mg COD/L 42%d 37%
d

85

References

Primary sludge

Food waste

Kitchen waste Organic fraction of Municipal solid waste

Batch reactor, pH 11, 60 C, 7 d, 0.02 g SDBSb/g VSS Batch reactor, pH 9, 35 C, 5 d Batch reactor, pH 8, 55 C, 9 d Batch reactor, 21 C, 6 d Batch reactor, 21 C, 6 d Batch reactor, 21 C, 6 d Batch reactor, pH 10, room temp., 5 d Continuous-ow completely mixed reactor, 25 C, HRT 1.25 d, SRT 10 d Semi-continuous reactor (once-a-day feeding and drawoff), pH 6, 35 C, HRT 8 d, OLR 9 g/L/d Batch reactor, 37 C, initial pH 5.5 Batch reactor, 35 C, 5 d, enzymatic pretreated food waste Batch reactor, pH 7, 35 C, 4 d Batch reactor, pH 45, 1422 C, HRT 44.5 d Plug ow reactor, pH 5.76.1, 37 C, HRT = SRT 6 d, OLR 38.5 g VS/L/d Plug ow reactor, pH 6.67.2, 55 C, HRT 6 d, OLR 22.4 g VS/L/d Batch reactor, pH 6.9, 40 C, cattail conc. 4.1 g VS/L Semi-continuous reactor (three times feeding per day), pH 6.5, 30 C, HRT 4 d Upow anaerobic sludge blanket reactor, pH 5.25.8, 35 C, HRT 0.9 d, OLR 16.6 g COD/L/d Batch reactor, initial pH 6.5, 25 C, 45 d Packed bed biolm reactor, pH 5.25.5, 25 C, HRT 1.4 d, OLR 26 g COD/L/d Continuous stirred-tank reactor, pH 5.5, 30 C, HRT 1.5 d, OLR 2.9 g COD/L/d Continuous stirred-tank reactor, pH 5.5, 30 C, HRT 1 d, OLR 6.5 g COD/L/d Continuous stirred-tank reactor, pH 6, 37 C, HRT 1 d Batch reactor, 1525 C, pH 6, 12 d Continuous-ow completely mixed reactor, 30 C, pH 6, HRT 0.67 d Continuous stirred-tank reactor, pH 6, 37 C, HRT 2.1 d Continuous ow-completely mixed reactor, pH 6.87.2, 35 C, HRT 0.5 d Upow anaerobic sludge blanket reactor, pH 5.5, 55 C, OLR 6 g COD/L/d Upow anaerobic sludge blanket reactor, pH 5.5, 37 C, HRT 0.5 d, SRT 15 d Upow anaerobic sludge blanket reactor, pH 6.5, 37 C, HRT 0.5 d Continuous-ow completely mixed reactor, pH 5.5, 35 C, HRT 0.5 d, OLR 13 g COD/L/d Batch reactor, 21 C, 6 d, mixing ratio 1:1 (on VSS basis) Semi-continuous reactor, 37 C, HRT = SRT 5 d Continuous-ow completely mixed reactor, 22 C, HRT 0.75 d, SRT 7 d, mixing ratio 1:1 (on volume basis) Continuous-ow completely mixed reactor, 25 C, HRT 1.25 d, SRT 10 d, mixing ratio 1:1 Semi-continuous reactor (once-a-day draw off and feeding), pH 6.99, 35 C, HRT = SRT 8.92 d, OLR 8.31 g VSS/ L/d, 88% food waste + 12% dewatered sludge (on VSS basis) Batch reactor, pH 8, 20 C, 4 d Continuous upow reactor, pH 5.55.9, 18 C, HRT 1 d, 25% food waste + 75% primary sludge (on weight basis) Semi-continuous reactor (daily feeding), pH 5.66.2, 35 C, HRT 2 d, mixing ratio 1:1 (on COD basis)

[2] [17] [17] [18] [14] [15] [16] [19] [89] [22] [23] [21] [24] [26] [25] [69] [27] [28] [30] [29] [88] [154] [10] [34] [35] [10] [31] [32] [33] [76] [73]

Wood mill efuent

0.75e 60%i 74%i 0.84e 3100 mg/L/d 1032 mg/L 2071 mg/L 1573 mg/L 44%d

Cheese whey Dairy wastewater

22,256 15,480 545c 392c

118 mg COD/g VSS 114 mg COD/g VSS 57 mg/g VSS/d 45 mg/g VSS/d 29,100 mg COD/L

[15] [155] [156] [19] [70]

Food waste + sludge

Not available

22,125 29,050 Sugar industry wastewater + pressed beet pulp


a b c d e f g h i

8237 mg COD/L 3610 mg/L 3635 mg/L (as acetic acid)

[157] [158] [159]

6621 (wastewater), 1220f (pulp)

mg TOC/L. Sodium dodecylbenzene sulfonate. sCOD after dilution. (mg VFA-COD in the efuent/mg COD in the inuent) 100%. (VFA-COD/wastewater inuent sCOD). mg COD/g dry weight. mg COD/kg. sCOD. (mg VFA-COD/mg sCOD) 100%.

86

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

VFA such as acetic, propionic and butyric acids. Both processes involve a complex consortium of obligate and facultative anaerobes, such as Bacteriocides, Clostridia, Bidobacteria, Streptococci and Enterobacteriaceae [12]. However, it is common practice that the hydrolysis and acidogenesis are conducted simultaneously in a single anaerobic reactor. In the past decades, numerous efforts had been devoted to maximize the production of VFA by exploring different types of wastes and regulating the operating conditions of the anaerobic reactor. It is now realized that proper process control can manipulate the type of VFA produced, which is critical to the performances of the downstream applications such as the production of polyhydroxyalkanoates (biodegradable plastics), electricity, biogas, and the biological removal of phosphorus and nitrogen from wastewater. To consolidate insights scattered across the literature, this article attempts to provide a review of the production and applications of waste-derived VFA with an outline as shown in Fig. 2. A systematic account of the different wastes used in VFA production is rst presented. Following that is the pretreatment methods used to enhance the hydrolysis of solid wastes. Anaerobic technologies for VFA production are introduced and the effects of main operational parameters of the anaerobic reactor, i.e., pH, temperature, retention time and organic loading rate on VFA production are discussed. Elucidation on the use of additives for enhancing VFA production is included. Finally, the various applications of wastederived VFA are discussed with the emphasis on the impact of the type of VFA on their performances.

2. Types of wastes for VFA production A variety of solid and liquid wastes, as presented in Table 1, have been studied for their potential to be used for VFA production. Among them, sludge, food waste and organic fraction of municipal solid waste (OFMSW) are the three most investigated solid wastes while wastewaters generated from the agricultural, dairy, pulp and paper industries are the liquid wastes frequently utilized for VFA production. Besides these, mixtures of different types of wastes had also been explored to enhance the production of VFA. Highlights about these various options are elaborated below. Primary sludge (PS) and waste activated sludge (WAS) generated from municipal wastewater treatment plant are frequently studied for VFA production because of the massive volumes generated from the widespread use of biological wastewater treatment [13]. Both PS and WAS are rich in organic matter with the total COD ranging from 14,800 mg/L to 23,000 mg/L [1418], making them promising wastes for the production of VFA. However, the soluble COD of the sludge is normally ten to hundred times less than its total COD [15]. This retards the production of VFA as the hydrolysis of the particulate organic matter in sludge is the ratelimiting step. Hence, more effort should be directed to improve the hydrolysis, e.g. through the application of pretreatment (Section 3). Instead of a single type of sludge, co-fermentation of a mixture of sludge had been proposed to improve the production of VFA. For instance, the maximum VFA production with PS alone was 85 mg COD/g VSS, but it increased by 40% to 118 mg COD/g VSS with the addition of WAS at a VSS ratio of 1:1 [15]. The enhanced VFA production was due to better synergistic hydrolysis of the sludge mixture [15]. Similarly, the addition of starch-rich industrial wastewater to PS increased the VFA production rate from 31 mg VFA/g VSS/day to 45 mg VFA/g VSS/day [19]. Food waste or kitchen waste is another solid waste commonly explored for VFA production for two reasons. First, food waste is a dominant component (2254%) in the huge volumes of municipal solid waste (MSW) [20]; and second, it has high total COD in a

range of 91,900166,180 mg/L [2123]. Nevertheless, one of the challenges is to effectively separate the food waste from MSW to minimize the interference of other components in the production of VFA. Source separation maybe a solution, but it requires high commitment from the public which is not easy to achieve. Another probable option is to establish a material recovery centre to separate the organic fraction of MSW (OFMSW) as well as glass, plastics, aluminum cans and ferrous metals. The OFMSW, with high total COD of 150,600347,000 mg/kg, is another possible source for VFA production [2426]. Similar to the case of using sludge for VFA production, the hydrolysis of OFMSW and food waste remains the rate-limiting step, thus calling for various pretreatment methods. For liquid waste, wastewaters generated from the agricultural [2730], dairy [10,3133], pulp and paper industries [10,34,35] are commonly used for VFA production. For instance, palm oil mill efuent is a potential agricultural wastewater with high COD of 88,000 mg/L and it can lead to the production of VFA with concentration up to 15,300 mg/L [27]. Cheese whey permeate and paper mill wastewater are also suitable due to its high readily fermentable organic content [10]. In contrast, the efuent from petrochemical industry is deemed unsuitable for VFA production regardless of its high COD value at 11,500 mg/L [36], owing to the presence of toxic and recalcitrant petrochemical pollutants which are harmful to the microorganisms. Compared with industrial wastewater, domestic wastewater has low organic content, with typical COD in the range of 175600 mg/L [37], thus it is not attractive for VFA production. As a whole it is still unclear which type of waste is more suitable for VFA production due to the use of different operating conditions and VFA production performance evaluation criteria, as shown in Table 1. However, wastes commonly used for VFA production, in general, are rich in organic matter with COD greater than 4000 mg/L (based on the reported values in Table 1). This could serve as a preliminary guide for waste selection. Besides, the ammonium content of waste should be lower than 5000 mg/L to avoid inhibition of VFA production [33] though it is an essential nitrogen source for the growth of microorganisms. Apart from the waste characteristics, the availability and the amount of the waste generated have to be taken into consideration to ensure stable and continuous waste supply for the production of VFA [38].

3. Pretreatment of solid wastes for VFA production In the production of VFA from solid waste, hydrolysis is the rate-limiting step [39] because of the complex structure and composition of the waste. For instance, cell wall and extracellular polymeric substances in sludge impose physical and chemical barriers to the hydrolysis of intracellular organic matter [40]. Meanwhile, lignocellulosic materials, fats and proteins in food waste and OFMSW lower their biodegradation rate [41,42]. Hence, various pretreatment methods (Table 2) have been explored to enhance the solubilization of the solid waste. Chemical pretreatment is effective for improving the hydrolysis of solid waste. The reagents commonly used in chemical pretreatment are acid [22,43], alkali [22,42,4446], ozone [4749] and hydrogen peroxide [50,51]. The application of acid and alkaline pretreatments helps in two ways: (i) to enhance the solubilization of the extracellular polymeric substances in the sludge [43,44] and (ii) to break the cell walls, resulting in the release of intracellular organic matter [43,46]. Besides sludge, these two pretreatment methods also promote the solubilization of food waste [22]. A signicant downside is the need for corrosion-resistant equipment due to the extreme values of pH used [22,43]. On the other hand, pretreatment with ozone is another option. Ozone is a strong

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 Table 2 Effect of different pretreatment methods on the solubilization of solid waste. Pretreatment methods Single pretreatment Acid Alkaline Solid wastes Food waste WAS OFMSW Food waste WAS WAS WAS + PS Ozone OFMSW WAS WAS OFMSW WAS OFMSW OFMSW OFMSW Food waste WAS WAS WAS OFMSW Kitchen waste WAS Pretreatment conditions HCl, pH 3, 4 C, 24 h HCl, pH 1, 24 h Ca(OH)2, dosage 2.3 g/L, 6 h NaOH, pH 11, 4 C, 24 h NaOH, pH 12.5, 30 min NaOH, pH 12 KOH, pH 13, 1 h Ozone dosage 0.16 g/g TS Ozone dosage 0.16 g/g TS Ozone dosage 1 g/h, 10 g TS/L, pH 6.8, 27 C, 1 h Hydrogen peroxide dosage 0.66g/g TS Hydrogen peroxide dosage 1 g/g TS Mature compost, inoculum loading 2.5% (v/v), 24 h Aspergullus awamori, inoculum loading 2.5% (v/v), 24 h Activated sludge from municipal wastewater treatment plant, inoculum loading 2.5% (v/v), 24 h Enzyme carbohydrase:protease:lipase at 1:2:1, enzyme dosage 0.1% (v/v), 45 C, initial pH 6.5, 24 h Cellulomonas uda and C. biazotea, 30 C, retention time 3 d Enzyme Accelerase 1500, enzyme dosage 0.07 g/ g VS, 50 C, 72 h Enzyme amylase:protease at 3:1 (w/w), 50 C, 12 h Microwave (2450 MHz), 41 min, heating from room temp. to 175 C Microwave (2450 MHz), 20.6 min, 7.8 C/min, heating from 22 C to 175 C Microwave (2450 MHz, 600W), 16 min, 02 min: <100 C, 210 min: 148 C (average), 1016 min: 171 C (average) Microwave (800 W), 3.5 min, nal temp. 80 C Microwave (2450 MHz), 15 min, heating from 10 C to boiling Frequency 20 kHz, specic energy 40,000 kJ/kg initial total solids Frequency 20 kHz, specic energy 90,692 J/g initial TS, ultrasonic density 0.4 W/mL, 1 h Frequency 20 kHz, specic energy 79 kJ/g TS Frequency 21 kHz, ultrasonic energy density 0.26 W/mL, 1 h Frequency 20 kHz, specic energy 9350 kJ/kg initial TS Frequency 20 kHz, specic energy input 45,000 kJ/ kg TS Frequency 42 kHz, 25 C, 2 h 121 C, 1.5 atm, 1 h 70 C, 30 min 121 C, 1.5 atm, 30 min 190 C, 1 h 170175 C, 1 h Effect of pretreatment sCOD increased by 28% sCOD increased by 4 times sCOD increased by 1.4 times sCOD increased by 28% sCOD increased by 66 times sCOD/TCOD ratio increased by 5 times sCOD/TCOD ratio increased by 3.7 times sCOD increased by 56% COD solubilization of 22%a sCOD increased by 25.8 times sCOD increased by 1.4 times sCOD increased by 3.9 times sCOD increased by 51% sCOD increased by 28% sCOD increased by 10% sCOD generation of 54.1% sCOD increased by 2.9 times sCOD increased by 4 times VSS reduction of 68% sCOD increased by 1.5 times sCOD increased by 2.5 times sCOD increased by 40 times References [22] [43] [42] [22] [44] [45] [46] [47] [48] [49] [51] [50] [39] [39] [39] [52] [55] [53] [54] [51] [41] [44]

87

Hydrogen peroxide Biological

Microwave

Sewage sludge WAS Ultrasound OFMSW OFMSW + sewage sludge Food waste WAS WAS WAS + PS WAS Thermal Food waste Food waste WAS WAS Sewage sludge + kitchen waste + fruit & vegetable waste Sewage sludge WAS WAS Combined pretreatment Alkaline + ultrasound Ozone + ultrasound

sCOD increased by 3.1 times sCOD increased by 14.5 times sCOD increased by 17.5% sCOD variation of 71.8% sCOD increased by 25% sCOD increased 8.7 times COD solubilization of 15%a sCOD/TCOD ratio increased by 5.7 times sCOD/TCOD ratio increased by 2.3 times sCOD generation of 17.9% sCOD increased by 20% sCOD/TCOD ratio increased by 2.2 times COD solubilization of 48%a VDS/VS ratio increased by 2.5 times sCOD increased by 25.6 times FVS/TVS ratio increased by 914%b FVS/TVS ratio increased by 751%b Disintegration degree of 69.4%c sCOD increased by 36.7 times

[56] [57] [47] [61] [22] [49] [48] [46] [45] [52] [22] [45] [48] [59]

90 C, 1 h 134 C, 1.5 h 70 C, 9 h

[60] [58] [58]

WAS + PS WAS

Thermochemical Thermochemical Alkaline + microwave

OFMSW WAS WAS

Alkaline = pH 13; Ultrasound = frequency 20 kHz, specic energy input 15,000 kJ/kg TS Ozone = ozone dosage 1 g/h; Ultrasound = frequency 21 kHz, ultrasonic energy density 0.26 W/mL 10 g TS/L, pH 6.8, 27 C, 1 h 180 C, 5 bar, 3 g NaOH/L, inert atmosphere (N2), 30 min 121 C, 7 g NaOH/L Alkaline = NaOH, pH 12.5, 30 min;

[46] [49]

sCOD increased by 224.5% sCOD increased by 10.8 times sCOD increased by 84 times

[39] [55] [44]

(continued on next page)

88 Table 2 (continued) Pretreatment methods Solid wastes

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

Pretreatment conditions Microwave = Microwave (2450 MHz, 600 W), 16 min, 02 min: <100 C, 210 min: 148 C (average), 1016 min: 171 C (average) Thermal = 121 C, 1.5 atm, 2 h; Biological = Enzyme carbohydrase:protease:lipase at 1:2:1, enzyme dosage 0.1% (v/v), 45 C, initial pH 6.5, 24 h Ultrasound = frequency 20 kHz, specic energy 79 kJ/g TS; Alkaline = NaOH, pH 11, 4 C, 24 h Ultrasound = frequency 20 kHz, specic energy 79 kJ/g TS; Acid = HCl, pH 3, 4 C, 24 h Ultrasound = frequency 20 kHz, specic energy 79 kJ/g TS; Thermal = 70 C, 30 min Hydrogen peroxide = Hydrogen peroxide dosage 0.66 g/g TS, room temp., 1 h; Microwave = Microwave (2450 MHz), 41 min, heating from room temp. to 85 C Hydrogen peroxide = Hydrogen peroxide dosage 1 g/g TS; Microwave = Microwave (2450 MHz), ramp time 5 min, ramping rate 20 C/min, heating time 5 min, nal temp. 120 C

Effect of pretreatment

References

Thermal + biological

Food waste

VSS reduction of 61.8%

[52]

Ultrasound + alkaline Ultrasound + acid Ultrasound + thermal Hydrogen peroxide + microwave

Food waste Food waste Food waste OFMSW

sCOD increased by 33% sCOD increased by 29% sCOD increased by 27% sCOD increased by 1.67 times

[22] [22] [22] [51]

Hydrogen peroxide + microwave

WAS

sCOD increased by 5 times

[50]

a b c

[(COD of supernatant after pretreatment COD of supernatant before pretreatment)/(particulate COD before pretreatment)] 100%. [(FVS/TVS ratio after pretreatment FVS/TVS ratio before pretreatment)/(FVS/TVS ratio before pretreatment)] 100%. [(sCOD after pretreatment sCOD before pretreatment)/(TCOD before pretreatment sCOD before pretreatment)] 100%.

oxidizing agent that can decompose into hydroxyl radicals to react with solid waste [49], leading to solubilization. Nevertheless, the high cost of ozone generation is a deterrent [47]. Hydrogen peroxide might be a less costly alternative, but it is less reactive as its oxidation potential (at 1.8 V) is lower than that of ozone at 2.1 V [50]. Biological agents such as hydrolytic enzymes [23,5254], Cellulomonas uda, C. biazotea [55], Aspergullus awamori [39], mature compost [39] and activated sludge from municipal wastewater treatment plant [39] had also been used to improve the solubilization of solid waste. In the pretreatment with enzyme, dosing a mixture of enzymes had greater impact on the hydrolysis of solid waste than a single enzyme [23,54] because different enzymes hydrolyze different components in the waste. Nonetheless, pretreatment using enzymes and pure cultures can be costly, thus inexpensive materials like mature compost and activated sludge should be considered in larger scale applications. Another signicant factor is the generally longer duration of the pretreatment by biological processes compared to other physicalchemical methods. Microwave pretreatment is another option [41,44,51,56,57], which enhances the hydrolysis of solid wastes via thermal and athermal (non-thermal) effects. The thermal effect is attributed to heat generation via the rotation of the dipolar molecules (e.g. water) in the oscillating electromagnetic eld [40]. The athermal effect is caused by the alignment of the polarized side chains of macromolecules (e.g. protein) with the poles of electromagnetic eld, thus breaking the hydrogen bonds and destabilizes the structure of the molecules [57]. Microwave pretreatment is energy intensive and it can lead to the formation of refractory (recalcitrant) compounds, such as melanoidins and humic acid [51], which lower the biodegradability of the treated solid waste. As a more conventional alternative to microwave, thermal pretreatment in the range of 60180 C [58] is another potential pretreatment method [22,45,48,52,5860]. At higher temperatures, other than greater energy input, it could, akin to microwave, lead to the formation of recalcitrant soluble organics [60]. Conversely, pretreatment at a lower temperature of 70 C could induce the thermophilic hydrolytic bacteria [60] which improve the solubilization of solid waste. The application of ultrasound to enhance the solubilization of solid waste had been investigated by many researchers

[22,4549,61]. Ultrasonic pretreatment leads to the formation of cavitation bubbles; and the collapse of bubbles generates hydromechanical shear forces strong enough to disrupt the macromolecules in the waste [62]. The collapse also raises the temperature and pressure, which lead to the formation of reactive hydroxyl radicals and thermal destruction of the solid waste [62]. Similar to microwave pretreatment, ultrasonic pretreatment is an energyintensive method [46]. This cost should be justied against the magnitude of improved solubilization of solid waste and the subsequent yield in VFA. In the literature, different pretreatment methods had been combined to promote the synergistic hydrolysis of solid waste [22,39,44,46,4952,55]. For instance, combined alkaline and ultrasonic pretreatment resulted in better disintegration of sewage sludge [46]. Alkaline pretreatment weakened the microbial cell wall, thus making it more vulnerable to the shear forces generated from the ultrasonic pretreatment. Another pretreatment combination is ultrasound and ozone [49] in which ultrasound enhances the decomposition of ozone into hydroxyl radicals while microbubbles of ozone help to produce more acoustic cavitations by acting as the cavitation nuclei. The possibilities are numerous and have yet to be exhaustively studied and analyzed for cost-benet. In selecting any or a combination of pretreatment methods above, it is important to consider the amount of solid wastes to be treated, the associated capital and operating costs, plus the extent of improvement in solubilization. Therefore, conducting a techno-economic feasibility study would be useful to identify the most suitable pretreatment method. In the literature, several narrowly-focused techno-economic evaluations had been carried out (e.g. [63,64]). For instance, Dhar et al. [63] conducted the technoeconomic evaluation of ultrasonic pretreatment, thermal pretreatment and their combination only. Feasibility studies with a wider coverage of pretreatment methods are necessary to complete the picture and to identify key aspects that require further cost-benet improvement.

4. Anaerobic technologies for VFA production Two common technologies used in the anaerobic production of VFA from waste are attached growth and suspended growth [65]. These technologies have led to the development of different types

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

89

Porous packing material

Fluidized bed

(a)
To gas chamber Gas-liquidsolid separator Sludge blanket

(b)

Clarifier Settled biomass

(c)

(d)

Fig. 3. Types of reactor used for anaerobic production of VFA (a) packed bed reactor, (b) uidized bed reactor, (c) upow anaerobic sludge blanket reactor (UASB) and (d) continuous stirred-tank reactor (CSTR) with recycling of biomass.

of reactor. For instance, packed bed reactor (Fig. 3a) is operated based on attached growth technology in which the biomass grows and attaches on porous packing material, such as alumina-based ceramic cubes [29] and granular activated carbon [66], inserted in the reactor. This retains the biomass in the reactor, thereby alleviating biomass washout. However, packed bed reactor will be clogged by wastes containing high concentrations of suspended solids. To avoid clogging, uidized bed reactor (Fig. 3b) has been developed. In this reactor, the biomass grows attached to small solid medium such as sand that remains in suspension by the upward owing motion of the uid [67]. In contrast, suspended growth technology allows the biomass to grow freely in suspension. Examples of reactors operated based on suspended growth technology are the upow anaerobic sludge blanket (UASB) reactor (Fig. 3c) and the continuous stirred-tank reactor (CSTR, Fig. 3d) [65]. Successful operation of the UASB reactor relies on the formation of dense and readily settleable biomass called granules. These granules are retained in the reactor by sedimentation, forming a sludge blanket at the bottom of the reactor. A gasliquidsolid separator is typically incorporated in the reactor [67] to return the less dense granules rising to the top of the reactor back to the sludge blanket. The drawback of the UASB reactor is a longer start-up period if the inoculum is not yet granulated [68]. For example, it took 80 days for the occurrence of granulation in the acidogenic fermentation of palm oil mill efuent [28]. On the other hand, the operation of CSTR involves complete mixing of waste and biomass. This can be approximately achieved by well-designed impellers, bafes and reactor shape. An associated challenge is to avoid using too high a speed of agitation such that the suspended microbes are damaged by the shear stress. When designed and operated properly, a CSTR is ideal to mix waste and microbes thoroughly in the presence of suspended solids in the waste. Typically, a gravity settling clarier is used to separate and to recycle the biomass from the efuent of the CSTR. The packed bed, uidized bed, UASB reactors and the CSTR are usually operated in the continuous mode; therefore they might not be feasible for reactions requiring long retention times (up to several days). For such slow reactions, some of these reactors can be converted into batch and semi-continuous reactors. For example, instead of feeding or withdrawing the contents of a CSTR continuously, both could be done intermittently, and the resulting reactor is a fed-batch reactor.

5. Factors inuencing VFA production The operational pH, temperature, retention time, organic loading rate, as well as additives used have great effects on the concentration, the yield and the composition of VFA produced from waste. In the literature, most of the researchers examine these factors one at a time and there are only a few works [69,70] evaluating their interactive effects. In view of that, the factors affecting VFA production will be discussed individually. 5.1. pH The pH value in the reactor is important to the production of VFA because most of the acidogens cannot survive in extremely acidic (pH 3) or alkaline (pH 12) environments [71]. The optimal pH values for the production of VFA are mainly in the range of 5.2511, but the specic ranges are dependent on the type of waste used (Table 3). For example, when sludge is used, the optimal pH values are in the range of 811. The alkaline condition enhances the hydrolysis of sludge through ionization of the charged groups (e.g. carboxylic groups) of the extracellular polymeric substances in the sludge [72], which are mainly carbohydrate and protein. This causes a strong repulsion between the extracellular polymeric substances [72], resulting in the release of carbohydrate and protein to the environment. Consequently, more soluble substrates are available for the production of VFA under alkaline conditions [16,17]. This is similar to the effect of alkaline pretreatment on the solubilization of sludge, as discussed in Section 3. Besides, the alkaline environment is not conducive to methanogenesis, thus preventing the consumption of the produced VFA for methane formation [17]. On the other hand, pH 7 was considered optimal for the hydrolysis and acidogenesis of kitchen waste as it led to the highest solubilization percentage of carbohydrate, protein and lipid as well as the highest VFA concentration in comparison with pH 5, 9 and 11 [21]. In contrast, the VFA production from wastewater is mostly conducted under acidic condition with optimum pH ranges from 5.25 to 6.0 [10,73]. Based on these ndings, it seems that alkaline condition favors the production of VFA from sludge whereas neutral and acidic conditions encourage the production of VFA from food waste and wastewater, respectively. In addition, pH can also affect the type of VFA produced from acidogenic fermentation, particularly acetic, propionic and butyric

90 Table 3 Optimal pH for the production of VFA. Type of wastes Primary sludge Waste activated sludge pH range studied 311 411 812 811 511 3.56 4.96 56.3

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

Optimal pH (range) 10 9 8 11 11 7 5.255.5 5.56 5.5

Reactor type and operating conditions Batch reactor, room temp., 5 d Batch reactor, 35 C, 5 d Batch reactor, 55 C, 9 d Batch reactor, 25 C, 4 d Batch reactor, 60 C, 7 d, 0.02 g SDBSa/g VSS Batch reactor, 35 C, 4 d Continuous stirred-tank reactor, 37 C, HRT 2 d Continuous stirred-tank reactor, 37 C, HRT 2 d Continuous-ow completely mixed reactor, 35 C, HRT 0.5 d, OLR 13 g COD/L/d

VFA production performance 60 mg COD/g VSS/d 298 mg COD/g VSS 368 mg COD/g VSS 1558 mg COD/L 259 mg TOC/g VSS 36,000 mg/L 0.83b 0.76b 44%c

References [16] [17] [160] [2] [21] [10] [73]

Kitchen waste Cheese whey Paper mill efuent Pharmaceutical wastewater


a b c

Sodium dodecylbenzene sulfonate. (VFA-COD/wastewater inuent soluble COD). (mg VFA-COD in the efuent/mg COD in the inuent) 100%.

acids [10,7477]. The production of propionic acid from dairy wastewater is favored at pH 44.5 whereas the acetic and butyric acids are favored at pH 66.5 [77]. Similar observation was reported in acidogenesis of gelatin-rich wastewater [76]. On the contrary, when pH increased from 5.25 to 6 in a study using cheese whey [10], the production of propionic acid increased while the production of acetic and butyric acids decreased. In another study, as pH increased from 6 to 8, the main VFA produced from glucoserich medium changed from butyric acid to acetic and propionic acids, and vice versa [74,75]. This might be caused by the shift in the dominant microbial populations from Clostridium butyricum at pH 6 to Propionibacterium at pH 8 [75]. The research ndings so far suggest that the optimal pH for the production of a specic VFA is highly dependent on the type of waste used. 5.2. Temperature The production of VFA from waste had been carried out under different temperature ranges, viz. psychrophilic (420 C) [19,78 80], mesophilic (2050 C) [2,19,76,7883], thermophilic (5060 C) [2,76,80,8284] and extreme/hyper-thermophilic (60 80 C) [8386] conditions. Increasing the temperature within the psychrophilic and mesophilic temperature ranges is benecial as it increases the concentration of VFA produced [78,79], the rate of VFA production [19] and the VFA yield [81]. For instance, raising temperature from 10 C to 35 C increased the VFA concentration produced from WAS by 300% [79]. The increment was due to the presence of greater amount of soluble carbohydrate and protein, which was a result of the improved sludge hydrolysis at higher temperature [79]. Similarly, the VFA production rate increased sixfold as the temperature increased from 8 to 25 C during the fermentation of PS [19]. Further increase in the operating temperature from mesophilic region to thermophilic region and to extreme/hyper-thermophilic region might still improve the VFA production. It had been reported [2] that thermophilic temperature (60 C) could lead to faster biological acclimatization and more active acidogenesis as compared to those at mesophilic temperature (35 C), thus leading to a higher VFA yield. Meanwhile, the production of VFA at extreme/hyper thermophilic temperatures of 7080 C had been found to outperform those at thermophilic temperatures of 55 60 C [84]. Nonetheless, the study of Yu et al. [83] showed that temperatures in the range of 4570 C had no effect on the production of VFA. On the contrary, Zhuo et al. [80] found that the acidforming enzymes activities at thermophilic temperature (55 C) were lower than that at mesophilic temperature (37 C). As a consequence, the total VFA concentration achieved at 55 C was 40%

lower than that at 37 C [80]. These inconsistent ndings were likely caused by the presence of different microbial species in the studies. If the production of VFA is more favorable at thermophilic or extreme/hyper-thermophilic temperatures, consideration must be given to the trade-off between the magnitude of improved VFA production and the heat requirement to maintain the temperature. A case in point is the acidogenesis of dairy wastewater, which was recommended to be carried out at mesophilic temperature in view of a lower energy demand and a more stable operation, in spite of the slightly higher production rate at thermophilic temperature [82]. The same recommendation was given to the acidogenesis of gelatin-rich wastewater [76]. Unlike pH, the inuence of temperature on the type of VFA produced is minor. Yuan et al. [78] performed the fermentation of WAS in mixed reactors at 4 C, 14 C and 24.6 C. As the temperature increased from 4 C to 14 C, the percentage of acetate reduced from 55% to 43%, but the percentage of propionate and butyrate increased slightly from 20% to 29% and from 11% to 16%, respectively. However, further increase in temperature to 24.6 C did not alter much the VFA composition. Similarly, no signicant variation in VFA composition was observed during the acidogenesis of gelatin-rich wastewater from 20 C to 55 C [76] and in the fermentation of ultrasonic-pretreated WAS from 10 C to 55 C [80]. Likewise, the composition of the VFA produced from synthetic dairy wastewater at 37 C and 55 C was comparable [82]. In the acidogenic fermentation of cellulose, acetic acid was the primary VFA produced at 37 C, 55 C and 80 C and butyric acid was the second dominant VFA [87]. Propionic acid was detected at 37 C only but its fraction in the VFA was relatively minor. Based on these results, the effect of temperature on VFA composition seems minor. This observation challenges the common understanding that microbial composition changes with temperature, and has raised the concern whether similar microbial species are present at different temperatures or different microbial species exist but produce similar types of VFA. To further understand the underlying reasons behind these observations, it would be interesting to examine the microbial community involved in the acidogenic fermentation at different temperatures. 5.3. Retention time In acidogenic fermentation of waste for the production of VFA, the retention time of the waste and the mixed microbial cultures in the anaerobic reactor are critical operational parameters. The former is termed hydraulic retention time (HRT), and the latter solids retention time (SRT). HRT is closely related to the initial capital

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

91

investment as it determines the volume of the reactor. On the other hand, SRT governs the selection of predominant microbial species in the reactor. Their inuences on VFA production will be discussed below. 5.3.1. Hydraulic retention time Applying higher hydraulic retention time (HRT) could be advantageous to the production of VFA [10,25,88] as the microorganisms have more time to react with the waste. For example, the production of VFA from OFMSW increased with HRT in a range of 26 d [25]. However, prolonged HRT could lead to stagnant VFA production [89,90]. For instance, the production of VFA from dairy wastewater nearly doubled as the HRT increased from 4 h to 12 h, but further increase to 1624 h only improved the VFA production by 6% [90]. Likewise, the VFA yield and volumetric productivity achieved in acidogenic fermentation of food waste increased as the HRT increased from 96 h to 192 h, but no signicant improvement was observed at a HRT of 288 h [89]. In the co-fermentation of WAS and fruit/vegetable waste, the concentration of VFA increased from 4400 mg/L to 6100 mg/L as the HRT increased from 1 d to 2 d [91]. Nonetheless, there was no signicant improvement on the production of VFA at higher HRT of 3 d (6150 mg/L) and 4 d (6620 mg/L) [91]. Other than performance considerations, operation with higher HRT requires a larger reactor and hence a greater cost [31]. In view of that, it is critical to determine the suitable HRT for any viable VFA production. Similar to pH, it was realized that HRT can be used to govern the relative production of propionic and butyric acids from paper mill efuent and whey [10]. During the acidogenic fermentation of whey, increasing the HRT from 20 h to 95 h favored the production of propionic acid but it suppressed the formation of butyric acid. Likewise, increasing HRT from 11 h to 24 h also promoted the production of propionic acid from paper mill efuent, but it disfavored the butyric acid production. However, HRT, in a range of 14 d, did not have signicant impact on the composition of VFA in the cofermentation of WAS and fruit/vegetable waste [91]. As HRT has the most signicant impact on the capital outlay, there is a strong need to ne-tune HRT to balance the distribution of VFA produced against the attainable total concentration. 5.3.2. Solids retention time In the case of using sludge for VFA production, SRT is equal to HRT because both the waste substrate and microbial mixed cultures are present in same phase. Most of the studies found that lower SRT is benecial to the production of VFA from sludge [9294]. This is because lower SRT can prevent the dominance of methanogens in the anaerobic reactor as the growth rate of methanogens is lower than that of acidogens [95]. Nevertheless, the SRT should be sufciently long to promote hydrolysis of the sludge. A case that illustrates this is the acidogenic fermentation of WAS, in which increasing the SRT from 4 d to 12 d led to 44% higher VFA concentration [93] because more soluble substrates resulting from sludge hydrolysis were available. However, further increase in SRT to 16 d resulted in lower VFA concentration although there were even more soluble substrates. It was deduced that the produced VFA was consumed by methanogens. Similarly, it was found that the acidogenic conditions prevailed at SRT 6 8 d whereas methanogenic condition prevailed at SRT P 10 d during the fermentation of PS [94]. However, there is very limited number of study on the inuence of SRT on the VFA production from wastewater [96]. Salmiati et al. [96] found that the concentration of VFA produced from palm oil mill efuent increased considerably for SRT of 6 d but reduced for 7 d. Meanwhile, other studies [33,77,82] normally applied SRT of 15 d in the acidogenic fermentation of wastewater.

The inuence of SRT on the VFA composition varies substantially from one study to another. The study of Feng et al. [93] showed that SRT, in the range of 416 d, had more inuence on the fraction of acetic and propionic acids relative to other higher molecular weight VFA in the acidogenic fermentation of WAS. Increasing SRT from 4 to 16 d caused the percentage of acetic acid to increase from 32% to 42% but the percentage of propionic acid to decrease from 24% to 14%. However, in another study using WAS [97], the percentage of acetic acid decreased instead from 66% to 49% with the increase of SRT from 5 d to 10 d whereas the percentage of propionic acid remained nearly constant at 1618%. The ndings were contradictory and it was probably due to the different operating conditions applied in these studies. The paucity of specic studies on the effects of SRT precludes rigorous conclusions on this point. 5.4. Organic loading rate Organic loading rate (OLR) shows the amount of waste, which can be expressed in terms of COD, VS, VSS or DOC, fed into the reactor daily per unit reactor volume. In the literature, the inuence of OLR on VFA production seemed inconsistent but could be rationalized by the presence of an optimum. For example, the VFA concentration produced from starchy wastewater increased linearly with OLR ranging from 1 g COD/L/d to 32 g COD/L/d [98]. Nevertheless, during the acidogenic fermentation of chemical synthesis-based pharmaceutical wastewater [73], the VFA concentration increased with OLR only in the range of 713 g COD/L/d. Worse, a slight increase to 14 g COD/L/d caused a drastic drop in the VFA concentration from 3410 mg/L (as acetate) to 1370 mg/L (as acetate) [73]. In a fermentation study of two-phase olive oil mill solid residue over an OLR range of 3.215.1 g COD/L/d, the maximum VFA concentration was achieved instead at an intermediate value of 12.9 g COD/L/d [99]. The VFA concentration produced from food waste [89] increased with OLR from 5 g/L/d to 13 g/L/d, but the operation of fermentor at 13 g/L/d was unstable because the fermentation broth became very viscous at high loading. These ndings suggest that different snapshots of bigger pictures were being observed. The linear dependence range [98] could be interpreted as behavior before the optimum, while the performance deteriorations [73,99] had crossed the optimum. In a case [89], rheology and the associated mass transfer implications appeared to be a signicant limiting factor outside the traditional operating parameters. Elucidation and moderation of the detrimental effects at high OLR will enable higher rates and intensity of waste treatment, further enhancing the economic feasibility of VFA production. Apart from the amount of waste fed into the anaerobic reactor, the frequency of feeding can also affect the production of VFA under semi-continuous conditions. The study of Nebot et al. [100] found that lower feeding frequency at 3 times per day can lead to a higher concentration of VFA as compared to that at 24 times per day. Less frequent feeding also offers easier reactor operation and reduces the wear and tear on the feed pumps. The OLR applied in the acidogenic fermentation has signicant inuence on the distribution of the VFA. In the fermentation of synthetic dairy wastewater [82], as the OLR increased from 4 g COD/L/d to 24 g COD/L/d, the percentage of acetate declined from 53% to 22% whereas the propionate percentage rose from 13% to 41% under mesophilic condition. Similar trend was observed in the thermophilic operation whereby the percentages of acetate and propionate changed from 44% to 23%, and from 21% to 43%, respectively. In another study on starchy wastewater [98], at a medium OLR of 10 g COD/L/d, propionic acid was the second main acid, but it was substituted by butyric acid at a higher OLR of 26 g COD/L/d. Throughout these ranges, acetic acid remained the primary VFA. The effects of OLR should be seen in the light that

92

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

Table 4 Effect of adding surfactants and/or enzymes on the production of VFA. Additive Waste T (C) Duration Additive dosage (day) 21 6 6 6 4 6 5 7 Nil 0.02 g/g dry sludge Nil 0.02 g/g TSS Nil 0.05 g/g dry sludge 0.1 g/g dry sludge 0.06 g/g dry sludge (amylase:protease = 1:3) SDS = 0.1 g/g dry sludge Enzyme = 0.06 g/g dry sludge (amylase:protease = 1:3) Maximum VFA References concentration (mg COD/L) 339 2599 118a 174a 191 1001 1143 1281 1457 [18] [15] [14] [102] [102] [102]

SDBS

Waste activated sludge

Waste activated sludge + primary sludge 21 SDS Waste activated sludge 21 50 50 50

a-Amylase + neuter protease Waste activated sludge SDS + a-amylase + neuter Waste activated sludge
protease
a

mg COD/g VSS.

there are a few other factors, notably pH and HRT, which strongly affect the distribution of the VFA. Incorporating these in a systematic investigation would help to clarify the specic effects of each parameter.

5.5. Additives In recent years, additives such as surfactants and enzymes have been utilized to improve the production of VFA from sludge, as summarized in Table 4. While some additives have hydrolytic/solubilizing effects on the waste, they are not considered as pretreatment agents because they are added during (instead of before) the acidogenic fermentation. Surfactants are known for their good solubilization ability. Thus, the addition of surfactants to sludge enhances the solubilization of the extracellular polymeric substances [14,18] which are mainly carbohydrate and protein [101]. This also breaks the sludge matrix resulting in the release of more inner carbohydrate and protein [14]. These, in turn, provide more substrates for the hydrolysis [13]. Sodium dodecyl sulfate (SDS) [14] and sodium dodecylbenzene sulfonate (SDBS) [18] are the surfactants commonly used to boost the production of VFA. With the addition of SDS at 0.05 g/g dry sludge, the maximum VFA concentration produced from WAS was 5 times higher than that in the absence of SDS [14]. It was observed that the maximum VFA concentration increased with the dosage of SDS ranging from 0.02 g/g dry sludge to 0.3 g/g dry sludge, but longer time was required to achieve maximal VFA production at higher SDS dosage [14]. This might be due to the stronger toxic effect of surfactant to acidogenic bacteria at higher dosages. On the other hand, the effect of using SDBS to enhance the VFA production was more dramatic. In a study using WAS [18], the maximum VFA concentration shot up nearly 8-fold from 339 mg COD/L to 2599 mg COD/L when 0.02 g/g dry sludge of SDBS was applied in the fermentation. Similar to SDS, longer time was required to reach the maximum VFA concentration at higher SDBS dosages [18]. Not limited to a single type of sludge, the addition of SDBS also positively affected the production of VFA from co-fermentation of WAS and PS [15]. The combined use of surfactant and hydrolytic enzymes (a-amylase and neuter protease) had been explored to improve the production of VFA. The results showed that the SDS-a-amylase-neuter protease system outperformed their individual systems in terms of sludge hydrolysis and VFA production [102]. The enhanced VFA production seen above has to be justied against the cost of the additives. In view of this, sludge with inherently high surfactant content should be exploited. Such potential sludge sources are from the wastewater treatment plants of the

textile nishing industry [13] and daily-use products [18]. A similar situation does not hold for wastewater or sludge inherently rich in the requisite enzymes, as these are rare if non-existent. Thus, the high cost of enzymes and the difcult enzyme recycling remain the main challenges in using enzymes to improve VFA production from waste [102]. Immobilization of the enzyme(s) might be an alternative to overcome this challenge. Functioning differently from surfactants and enzymes, chemical inhibitors of methanogenesis might enhance the production of VFA by suppressing the activity of VFA-consuming methanogens. Recently, Liu et al. [103] had reviewed the types of chemical methanogenic inhibitors and their respective inhibition mechanisms. In brief, the chemical inhibitors of methanogenesis can be categorized as specic and nonspecic inhibitors. The former inhibits the specic enzymes only existing in methanogens such as coenzyme M (inhibitors: 2-bromoethanesulfonate and 2-chloroethanesulfonate) and hydroxymethylglutary-SCoA reductase (inhibitors: mevastatin and lovastatin). On the other hand, the nonspecic inhibitors affect the activity of both methanogens and non-methanogens. Examples of nonspecic inhibitor are ethylene, acetylene and several halogenated aliphatic hydrocarbons like chloroform, uoroacetate and methyl uoride. For VFA production, it is necessary to evaluate and clarify the effects (if any) of these inhibitors, particularly the non-specic ones, on the viability of acidogens. 6. Applications of waste-derived VFA The VFA produced from acidogenic fermentation of waste is a valuable substrate to a variety of applications such as the production of biodegradable plastics, generation of bioenergy and biological nutrient removal. It is often feasible, and certainly highly desirable, to utilize the VFA-rich fermented waste directly in these applications. However, for some applications, the fermented waste has to be further treated to enhance the performance. The treatment involved would be discussed where relevant. In the following, to provide guidance on the optimization targets of the acidogenic fermentation that generates the VFA, the main focus will be on the impact of the type of VFA on the performance of the aforementioned applications. 6.1. Polyhydroxyalkanoates Polyhydroxyalkanoates (PHA) are biodegradable polymers that can be synthesized by microorganisms using renewable resources such as VFA. Although PHA has a broad range of applications in various industries [104] and is environmental-friendly, its substitution over the conventional petrochemical-based plastic is limited

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

93

by the high production cost. This is attributed to the use of expensive and well-dened carbon substrate which contributed to about 31% of the total operating cost [105]. In view of that, low-cost substrate such as waste-derived VFA is a promising option. Before the fermented waste that is rich in VFA can be utilized for PHA production, it is important to regulate its ammonium and phosphorus contents because excessive nutrients would favor the growth of microorganisms and reduce the conversion of VFA to PHA [106]. Further, it has been reported that limited nitrogen and phosphorus conditions could lead to higher PHA content and yield [35]. Excessive nitrogen and phosphorus, if present in the fermented waste, can be removed simultaneously via struvite precipitation. This technique had been proven to be effective for rapid ammonium and phosphorus removal from fermented WAS with negligible VFA loss [2]. In addition, direct feeding of fermented waste with considerable amount of sludge particles into the PHA production reactor is not advisable as it can lead to the failure of PHA production [107]. In general, the fermented waste should be ltered before use [98,106]. Combined ltration and evaporation system can be considered, if it is desirable to use pure VFA in the production of PHA [108]. Nevertheless, pure VFA is rarely needed except in some cases of PHA production by pure microbial culture [27]. The PHA content achieved by pure microbial culture is higher generally, as shown in Table 5, but sterile condition is required and this would defeat the main purpose of reducing the PHA production cost as it involves additional energy input and equipment. On the contrary, PHA production by mixed microbial culture eliminates the need of sterilization and hence it is a more economically feasible production route for waste-derived VFA. The PHA content obtained from mixed culture can be improved by optimizing the operational conditions of the cultivation reactor of PHA accumulating organisms [34,109,110], by feeding the suitable VFA type [111] and/or by ne-tuning the PHA production conditions [112114]. With these strategies, PHA content in a range of 4077% can be achieved by mixed microbial cultures fed with fermented food waste [114], fermented waste activated sludge [113], fermented sugar cane molasses [110] and fermented paper mill efuent [34]. The chain-length of the VFA has great inuence on the composition and hence the mechanical properties and application of the resulting PHA. In mixed culture PHA production, acetic and butyric acids favor the production of 3-hydroxybutyrate (3HB) whereas propionic and valeric acids promote the synthesis of 3-hydroxyvalerate (3HV) [106,111,115117]. Poly(3-hydroxybutyrate) (P(3HB)) is brittle and stiff, and thus has limited applications. The incorpo-

ration of 3HV into P(3HB) leads to the formation of copolymer P(3HB-co-3HV) which is more exible and tougher [118]. Besides, it is less permeable to oxygen as compared to the commercial polyethylene and polypropylene, making it a suitable food packaging material [38]. Hence it is essential to regulate the VFA composition during acidogenic fermentation to facilitate the production of PHA with desired properties. 6.2. Bioenergy The increase in oil prices and the depletion of oil reserve is brewing an energy crisis, which is one of the biggest challenges in the 21st century. Therefore, it is essential to develop alternative energy generation route and to diversify the energy source available in the current market. Waste-derived VFA is an inexpensive energy source that can be used for the generation of different forms of energy. Direct electricity generation from waste-derived VFA is possible through the use of microbial fuel cell. Furthermore, it can be used to produce various valuable fuels such as biogas, hydrogen and biodiesel. 6.2.1. Electricity Microbial fuel cell (MFC) is a bioelectrochemical system that uses microorganisms to harness the chemical energy of the organic substrate as a source of electricity. Fig. 4 helps to illustrate the working principles of a MFC with VFA as the model organic substrate. The MFC consists of an anaerobic anodic chamber and an aerobic cathodic chamber which are separated by a proton exchange membrane [119]. At the anode, a biolm is formed on which microorganisms oxidize the VFA to generate electrons, protons and carbon dioxide. The protons pass through the proton exchange membrane to enter cathodic chamber whereas the electrons ow through an external circuit to the cathode. At the cathode, the electrons combine with protons and oxygen to produce water. This completes the electrical circuit and drives a current. The anodic and cathodic reactions are shown in Eqs. (1) and (2), respectively, using acetate as an example organic substrate [119].

CH3 COO 2H2 O ! 2CO2 7H 8e O2 4e 4H ! 2H2 O

1 2

Several types of MFCs had been used for electricity generation such as the two-compartment MFC (Fig. 4), the single-compartment MFC, the upow mode MFC and the stacked MFC. The design and the operation of these MFCs have been detailed in a review by Du et al. [119]. In most of the MFC operation, VFA-rich fermented

Table 5 PHA production by pure and mixed microbial cultures from waste-derived VFA. Source of VFA Fermented starchy wastewater Fermented palm oil mill efuent Pure culture Alcaligenes eutrophus Ralstonia eutropha Comamonas sp. EB 172 Mixed culture Activated sludge Activated sludge Activated sludge Activated sludge Activated sludge Activated sludge enriched with glycogen-accumulating organisms Activated sludge Activated sludge Activated sludge Activated sludge PHA contenta (wt%) 34 >90 86 40 48 77 29 75 37 57 73 40 48 References [98] [27] [161] [96] [35] [34] [88] [110] [115] [2] [113] [114] [162]

Fermented palm oil mill efuent Fermented paper mill wastewater Fermented wood mill efuent Fermented sugar cane molasses Fermented waste activated sludge Fermented food waste Fermented food waste and sewage sludge
a

(PHA/cell dry weight or VSS or TSS) 100%.

94

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

Resistor eeProton exchange membrane H2O H+ CO2 Anode Anaerobic chamber Biofilm Cathode Aerobic chamber O2

VFA

Fig. 4. Electricity generation from volatile fatty acids through a microbial fuel cell (adapted from [153]).

anaerobic digestion [128]. A notable feature of both one-phase anaerobic digestion and two-phase anaerobic digestion is that the VFA produced is used as it is without further treatment, making the process very cost-effective. Among all the VFA, methanogens seem to have the lowest tolerance for inhibition by propionic acid [129,130]. Complete inhibition triggered by propionate occurred at 5000 mg/L which was ve times lower than the inhibition concentration of acetate and butyrate [129]. Besides, the methanogenic bacteria concentration dropped from 6 107 1/mL to 1 107 1/mL as the concentration of propionic acid reached 900 mg/L [130]. However, there was no signicant inhibition at higher acetate and butyrate concentrations of 2400 mg/L and 1800 mg/L, respectively. Based on these research ndings, it is critical to maintain a low propionate concentration in the waste-derived VFA to avoid inhibition of methanogenesis. 6.2.3. Hydrogen Other than as a by-product of the two-phase anaerobic digestion above, hydrogen production from the waste-derived VFA can be achieved by photo fermentation [131], electrohydrolysis [132] or microbial electrolysis cell [133]. In photo fermentation, purple non-sulfur bacteria convert VFA into hydrogen in the presence of light [134]. The photo fermentation is frequently employed together with dark fermentation as a two-stage hydrogen production process [11,135,136]. This is because dark fermentation not only produces hydrogen but also VFA. The resulting VFA-rich efuent from dark fermentation can then be used in photo fermentation to enhance the overall hydrogen production. Prior to photo fermentation, it is necessary to remove the colloidal particles in fermented waste to ensure good light penetration [137]. Besides, the fermented waste should have low ammonium content to avoid the inhibition of nitrogenase enzyme which catalyzes the photo-hydrogen production [138]. Dilution has been proposed to reduce the ammonium concentration, but this could lead to insufcient nutrients [139] and the requirement of larger reactor [138]. Thus, the development of a more economical ammonia removal technique is required in future. The hydrogen production is also sensitive to the type of VFA applied in the photo fermentation process. Rhodobacter sphaeroides, a bacterial strain commonly used in photo fermentation, could better utilize acetate and propionate for hydrogen production as compared to butyrate [3]. The maximum hydrogen production rate from acetate and propionate was four times higher than that of butyrate. These indicate the need to increase the feed acetate and propionate concentrations for better hydrogen production performance by R. sphaeroides. Electrohydrolysis of VFA is another option for hydrogen production and its mechanism is detailed as follows. With the application of direct current voltage, electrons are released from the metal electrode (using copper electrode as an example) and combine with protons generated from electrohydrolysis of VFA to produce hydrogen [132]. The aforementioned reactions are presented in the following equations:

waste can be used directly for electricity generation without any treatment [4,120,121], thus making this technology economically attractive. The type of VFA used in MFC has considerable inuence on the performance of electricity generation performance (Table 6). The amount of current produced from acetate-fed MFC was approximately two times higher than the other higher molecular weight VFA while achieving the highest coulombic efciency (CE) of 93% [122]. Similarly, acetate-fed MFC also outperformed the propionate- and butyrate-fed MFC in terms of CE and power density (PD) [123]. Higher PD could be attained when acetate was the main VFA in a mixture of acetate, propionate and butyrate [124]. Several studies demonstrated that anodic microorganisms have preference over lower molecular weight VFA [4,122]. The consumption of acetate and propionate in a MFC fed with fermented food waste was faster than other higher molecular weight VFA [4]. Likewise, the extent of acetate and propionate consumption in the MFC fed with mixed VFA was higher, but valerate was scarcely removed after more than 10 weeks of operation [122]. These ndings imply the need to increase the fraction of acetate in the waste-derived VFA for better electricity generation. 6.2.2. Biogas Biogas is commonly used for heat and power generation because of its high methane content (6570 v/v%) [65]. The production of biogas is carried out under anaerobic conditions using VFA as the precursor. It is feasible to produce biogas from waste using single anaerobic digester with VFA as the intermediate product. This process is known as one-phase anaerobic digestion. However, the optimal growth-and-function conditions of the microorganisms responsible for acidogenesis and methanogenesis are vastly different. It is difcult to provide optimal conditions for both groups of microorganisms in a single digester, resulting in suboptimal performance of the one-phase anaerobic digestion [125]. This problem can be resolved through the application of twophase anaerobic digestion in which the two groups of microorganisms are divided into two separate digesters. The rst digester operates at acidic pH and short SRT to cultivate fast-growing acidogens whereas the second digester has neutral pH and higher SRT to enrich slow-growing methanogens [67]. The two-phase anaerobic digestion can operate at higher OLR and to achieve better biogas productivity as compared to one-phase anaerobic digestion [126]. Another attractive feature of two-phase anaerobic digestion is the simultaneous recovery of hydrogen and methane; the former from the rst digester while the latter from the second [127]. This feature would lead to a 25% higher energy yield than the one-phase

Cu ! Cu2 2e CH3 COOH ! CH3 COO H 2H 2e ! H2

3 4 5

An interesting feature of electrohydrolysis of VFA is that it can be incorporated into anaerobic reactor to allow in situ hydrogen production [140]. This permits direct utilization of waste-derived VFA while achieves space saving. However, to date, the number of reported studies on hydrogen production by electrohydrolysis of waste-derived VFA is scarce [132,140]. There is no study that we are aware of on the effect of the type of VFA on hydrogen production, or on less costly but effective sacricial electrodes.

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 Table 6 Performance of electricity generation from VFA via microbial fuel cell. Volatile fatty acids Electricity generation performance Current (mA) Acetate Propionate n-Butyrate i-Butyrate n-Valerate i-Valerate Hexanoate Mixture of acetate, propionate, butyrate and valerate Acetate Propionate Butyrate Mixture of acetate, propionate and butyrate Fermented food waste Leachate of fermented municipal solid waste
a

95

Electrodes Anode Granular graphite Cathode Granular graphite

References

Power density (mW/m2) 49a 64 58 51 1.32a 15 5.9a

Coulombic efciency (%) 93 55 49 67 33 34 44 72 36 43 38 6

28 13 16 5 12 13 12 21

[122]

Carbon felt

Perforated titanium plate coated with platinum

[123]

Carbon ber Carbon felt Carbon cloth

Carbon cloth coated with platinum Carbon paper coated with platinum Carbon cloth coated with platinum

[124] [4] [163]

W/m3.

Table 7 Nitrogen and phosphorus removal efciencies of BNR process using waste-derived VFA as carbon source. Source of VFA Fermented Fermented Fermented Fermented food waste waste activated sludge waste activated sludge waste activated sludge Nitrogen removal efciency (%) 92 83 82 Phosphorus removal efciency (%) 73 93 95 99 References [149] [164] [5] [148]

Another electrochemical option is the microbial electrolysis cell that produces hydrogen through cathodic reduction of proton released from the microbial oxidation of VFA at the anode [141]. Again, this process requires the supply of external electricity. The reactions involved at anode and cathode are shown in Eqs. (6) and (7), respectively [141].

CH3 COOH 2H2 O ! 2CO2 8e 8H 8H 8e ! 4H2


6 7

Synthetic acetate is normally used as the model VFA in hydrogen production by microbial electrolysis cell [141]. It was only in recent years that waste-derived VFA was found to be feasible for hydrogen production via a microbial electrolysis cell [133]. Moreover, it was observed that the consumption of acetate was greater than other VFA in the fermentation liquid generally [133]. This nding hints that the hydrogen production might be improved by increasing the fraction of acetate in feed as the anodic microorganisms prefer simpler VFA. 6.2.4. Lipids for biodiesel Biodiesel is usually a methyl ester of long-chain fatty acids that can be produced from lipids through transesterication process. Although biodiesel is a renewable energy resource, its commercialization has been prevented by the high production cost attributed to the use of expensive raw material which accounts for 7075% of the total cost [142]. Currently, edible lipid obtained from agricultural commodities such as rapeseed oil, palm oil and soybean oil are commonly used in biodiesel production [143]. This also raises the ethical concern of using food for fuel. Hence, the search for inexpensive non-edible lipid is underway. The microbial lipid synthesized from waste-derived VFA by oleaginous microorganisms

offers a sensible alternative [144]. Furthermore, the microbial lipid synthesized from VFA has similar fatty acid compositions to soybean oil and jatropha oil, thus making it suitable for biodiesel production [145]. To date, there is limited research work that explored the microbial lipid production from waste-derived VFA [144] as most of the studies used synthetic VFA instead [145,146]. The lipid content obtained from Cryptococcus curvatus fed with VFA derived from food waste was considerably low at 13.8% [144]. The lower lipid content was attributed to the low carbon-to-nitrogen ratio in the fermented food waste at 3.2:1 which led to unfavorable lipid production [144]. This nding implies the need to remove nitrogenous compounds in the fermented waste for achieving higher lipid content. Another alternative to increase the lipid content is through the manipulation of the feed VFA composition [145]. It was found that the feeding of VFA with a ratio of acetic acid: propionic acid: butyric acid of 8:1:1 to Cryptococcus albidus led to the highest lipid content (27.8%) as compared to ratios of 4:3:3 (19.8%), 6:1:3 (27.3%) and 7:2:1 (26.1%).

6.3. Biological nutrient removal VFA is an important carbon substrate to assist the biological removal of nitrogen and phosphorus from wastewater. It is known that nitrogen removal can be accomplished through the aerobic nitrication followed by anoxic denitrication. Meanwhile, phosphorus can be removed via enhanced biological phosphorus removal (EBPR) process conducted under alternating anaerobic and aerobic conditions. Simultaneous nitrogen and phosphorus removal is feasible through the application of alternate anaerobic aerobicanoxic conditions.

96

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399

Under most circumstances, additional carbon substrate such as VFA is required for stable biological nutrient removal (BNR) processes because the carbon substrate in wastewater is often insufcient. The carbon to nitrogen requirement should be practically in the range of 510 mg COD/mg N for combined nitrication/denitrication [147]. In addition, 7.510.7 mg of COD is required to remove 1 mg of phosphorus [67]. Using synthetic VFA can be expensive, thus a more economical solution is waste-derived VFA. Before a fermented waste is used, its nitrogen and phosphorus contents should be examined. Struvite precipitation can be employed to remove excessive amounts of nitrogen and phosphorus from fermented waste [148]. However, this might not be necessary as high nitrogen and phosphorus removal efciencies can still be achieved in BNR without the prior removal of nitrogen and phosphorus in the fermented food waste [149]. Several studies demonstrated that using waste-derived VFA resulted in superior BNR performance than using synthetic acetate [5,148], presumably due to the synergistic effects of other components present in the former. Higher phosphorus removal efciency was achieved with the use of VFA derived from WAS (98.7%) as compared to acetate (71.1%) [148]. The higher efciency was attributed to the lower glycogen synthesis and higher phosphorus uptake for every amount of PHA degraded. Zheng et al. [5] compared the nitrogen and phosphorus removal efciencies using VFA produced from WAS fermentation and synthetic acetate under alternate anaerobicaerobicanoxicaerobicanoxicaerobic conditions. The former again led to higher nitrogen and phosphorus removal efciencies at 82% and 95%, respectively. The authors explained that the presence of propionate was probably the reason for better phosphorus removal while the higher nitrogen removal efciency might be due to the better use of exogenous denitrication pathway for nitrogen removal. As summarized in Table 7, the high nitrogen and phosphorus removal efciencies associated with the use of waste-derived VFA afrm that it is an excellent, costeffective carbon source for BNR processes. Similar to other applications alluded to above, there is differential performance of BNR depending on the type of the VFA. For the nitrogen removal process, denitrifying bacteria have preference for lower molecular weight VFA. Acetate is normally the rst VFA to be consumed, followed by propionate and butyrate, and lastly valerate [150]. The order of VFA preference by denitrifying bacteria might be associated with the simpler metabolic pathway used in the assimilation of lower molecular weight VFA [150]. Besides, the average specic denitrication rate for acetate is two times higher than that of propionate [150] indicating the need to increase the acetate fraction if waste-derived VFA is the carbon source in denitrication. Unlike denitrication which prefers acetate, Chen et al. [151] reported that increasing propionic acid content in the domestic wastewater leads to higher phosphorus removal efciency in the long run. When the ratio of propionic acid to acetic acid increased from 0.16 mM carbon/mM carbon to 2.06 mM carbon/mM carbon, the phosphorus removal efciency improved from 77% to 87% contributed by the greater phosphorus uptake per unit PHA degraded or the lower fraction of glycogen accumulating organisms in the EBPR process.

regulate the composition of the VFA produced which is critical to downstream applications. In addition to controlling the classical operating conditions, the production of VFA can be enhanced through the application of various pretreatment methods and the addition of surfactants, hydrolytic enzymes and chemical methanogenic inhibitors, but the extra cost involved has to be justiable. On the other hand, it is also essential to examine the microbial community involved in the acidogenic fermentation. This can help to clarify how different operating conditions affect the microbial dynamics, which is useful in devising an optimal operating strategy. Seen macroscopically, studies that examine all pertinent parameters above via suitably designed experiments remain very few, hence the synergistic or antagonistic interactions amongst the variables remain unsatisfactorily examined. Such uncertainties in part contributed to the fact that most of the applications of waste-derived VFA still remain in the laboratories. To ascertain the transferability of the technologies from the laboratory to the commercial market, pilot-scale studies are also required, not only to ne-tune the technical aspects, but also to scrutinize the economics. When all these are accomplished, the winning candidate technologies in the production and application of waste-derived VFA would lead to a more environmental-friendly and sustainable society. Acknowledgment The authors would like to acknowledge the University of Malaya Postgraduate Research Grant (PV061/2011A) and the University of Malaya Research Grant (RP002C-13AET) for funding this work. References
[1] APHA, Standard Methods for the Examination of Water and Wastewater, 18th ed., APHA, United States of America, 1992. [2] M. Cai, H. Chua, Q. Zhao, N.S. Sin, J. Ren, Optimal production of polyhydroxyalkanoates (PHA) in activated sludge fed by volatile fatty acids (VFAs) generated from alkaline excess sludge fermentation, Bioresour. Technol. 100 (2009) 13991405. [3] B. Uyar, I. Eroglu, M. Ycel, U. Gndz, Photofermentative hydrogen production from volatile fatty acids present in dark fermentation efuents, Int. J. Hydrogen Energy 34 (2009) 45174523. [4] J. Choi, H.N. Chang, J. Han, Performance of microbial fuel cell with volatile fatty acids from food wastes, Biotechnol. Lett. 33 (2011) 705714. [5] X. Zheng, Y. Chen, C. Liu, Waste activated sludge alkaline fermentation liquid as carbon source for biological nutrients removal in anaerobic followed by alternating aerobicanoxic sequneching batch reactors, Chin. J. Chem. Eng. 18 (2010) 478485. [6] Y.L. Huang, Z. Wu, L. Zhang, C.M. Cheung, S. Yang, Production of carboxylic acids from hydrolyzed corn meal by immobilized cell fermentation in a brous-bed bioreactor, Bioresour. Technol. 82 (2002) 5159. [7] E. Akaraonye, T. Keshavarz, I. Roy, Production of polyhydroxyalkanoates: the future green materials of choice, J. Chem. Technol. Biotechnol. 85 (2010) 732 743. [8] J. Zigov, E. turdk, D. Vandk, . Schlosser, Butyric acid production by Clostridium butyricum with integrated extraction and pertraction, Process Biochem. 34 (1999) 835843. [9] T. Kondo, M. Kondo, Efcient production of acetic acid from glucose in a mixed culture of Zymomonas mobilis and Acetobacter sp, J. Ferment. Bioeng. 81 (1996) 4246. [10] S. Bengtsson, J. Hallquist, A. Werker, T. Welander, Acidogenic fermentation of industrial wastewaters: effects of chemostat retention time and pH on volatile fatty acids production, Biochem. Eng. J. 40 (2008) 492499. [11] H. Su, J. Cheng, J. Zhou, W. Song, K. Cen, Improving hydrogen production from cassava starch by combination of dark and photo fermentation, Int. J. Hydrogen Energy 34 (2009) 17801786. [12] P. Weiland, Biogas production: current state and perspectives, Appl. Microbiol. Biotechnol. 85 (2010) 849860. [13] S. Jiang, Y. Chen, Q. Zhou, Inuence of alkyl sulfates on waste activated sludge fermentation at ambient temperature, J. Hazard. Mater. 148 (2007) 110115. [14] S. Jiang, Y. Chen, Q. Zhou, Effect of sodium dodecyl sulfate on waste activated sludge hydrolysis and acidication, Chem. Eng. J. 132 (2007) 311317. [15] Z. Ji, G. Chen, Y. Chen, Effects of waste activated sludge and surfactant addition on primary sludge hydrolysis and short-chain fatty acids accumulation, Bioresour. Technol. 101 (2010) 34573462.

7. Conclusions This review has demonstrated that VFA can be produced from various wastes by acidogenic fermentation and it is a suitable carbon substrate for the production of bioplastics, bioenergy, and biological phosphorus and nitrogen removal. Proper control of the pH, temperature, HRT, SRT and OLR is important to ensure successful VFA production. More importantly, most of these parameters also

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 [16] H. Wu, D. Yang, Q. Zhou, Z. Song, The effect of pH on anaerobic fermentation of primary sludge at room temperature, J. Hazard. Mater. 172 (2009) 196 201. [17] P. Zhang, Y. Chen, Q. Zhou, Waste activated sludge hydrolysis and short-chain fatty acids accumulation under mesophilic and thermophilic conditions: effect of pH, Water Res. 43 (2009) 37353742. [18] S. Jiang, Y. Chen, Q. Zhou, G. Gu, Biological short-chain fatty acids (SCFAs) production from waste-activated sludge affected by surfactant, Water Res. 41 (2007) 31123120. [19] I. Maharaj, P. Elefsiniotis, The role of HRT and low temperature on the acidphase anaerobic digestion of municipal and industrial wastewaters, Bioresour. Technol. 76 (2001) 191197. [20] S. Kathirvale, M.N.M. Yunus, K. Sopian, A.H. Samsuddin, Energy potential from municipal solid waste in Malaysia, Renew. Energy 29 (2003) 559567. [21] B. Zhang, L.-L. Zhang, S-C. Zhang, H.-Z. Shi, W.-M. Cai, The inuence of pH on hydrolysis and acidogenesis of kitchen wastes in two-phase anaerobic digestion, Environ. Technol. 26 (2005) 329340. [22] E. Elbeshbishy, H. Hafez, B.R. Dhar, G. Nakhla, Single and combined effect of various pretreatment methods for biohydrogen production from food waste, Int. J. Hydrogen Energy 36 (2011) 1137911387. [23] H.J. Kim, S.H. Kim, Y.G. Choi, G.D. Kim, T.H. Chung, Effect of enzymatic pretreatment on acid fermentation of food waste, J. Chem. Technol. Biotechnol. 81 (2006) 974980. [24] D. Bolzonella, F. Fatone, P. Pavan, F. Cecchi, Anaerobic fermentation of organic municipal solid wastes for the production of soluble organic compounds, Ind. Eng. Chem. Res. 44 (2005) 34123418. [25] C. Sans, J. Mata-Alvarez, F. Cecchi, P. Pavan, A. Bassetti, Acidogenic fermentation of organic urban wastes in a plug-ow reactor under thermophilic conditions, Bioresour. Technol. 54 (1995) 105110. [26] C. Sans, J. Mata-Alvarez, F. Cecchi, P. Pavan, A. Bassetti, Volatile fatty acids production by mesophilic fermentation of mechanically-sorted urban organic wastes in a plug-ow reactor, Bioresour. Technol. 51 (1995) 8996. [27] S.K. Hong, Y. Shirai, A.R.N. Aini, M.A. Hassan, Semi-continuous and continuous anaerobic treatment of palm oil mill efuent for the production of organic acids and polyhydroxyalkanoates, Res. J. Environ. Sci. 3 (2009) 552 559. [28] R. Borja, C.J. Banks, E. Snchez, Anaerobic treatment of palm oil mill efuent in a two-stage up-ow anaerobic sludge blanket (UASB) system, J. Biotechnol. 45 (1996) 125135. [29] M. Beccari, L. Bertin, D. Dionisi, F. Fava, S. Lampis, M. Majone, F. Valentino, G. Vallini, M. Villano, Exploiting olive oil mill efuents as a renewable resource for production of biodegradable polymers through a combined anaerobic aerobic process, J. Chem. Technol. Biotechnol. 84 (2009) 901908. [30] D. Dionisi, G. Carucci, M.P. Papini, C. Riccardi, M. Majone, Olive oil mill efuents as a feedstock for production of biodegradable polymers, Water Res. 39 (2005) 20762084. [31] B. Demirel, O. Yenigun, Anaerobic acidogenesis of dairy wastewater: the effects of variations in hydraulic retention time with no pH control, J. Chem. Technol. Biotechnol. 79 (2004) 755760. [32] H.Q. Yu, H.H.P. Fang, Thermophilic acidication of dairy wastewater, Appl. Microbiol. Biotechnol. 54 (2000) 439444. [33] H.Q. Yu, H.H.P. Fang, Acidication of mid- and high-strength dairy wastewaters, Water Res. 35 (2001) 36973705. [34] Y. Jiang, L. Marang, J. Tamis, M.C.M. van Loosdrecht, H. Dijkman, R. Kleerebezem, Waste to resource: converting paper mill wastewater to bioplastic, Water Res. 46 (2012) 55175530. [35] S. Bengtsson, A. Werker, M. Christensson, T. Welander, Production of polyhydroxyalkanoates by activated sludge treating a paper mill wastewater, Bioresour. Technol. 99 (2008) 509516. [36] P. Ghosh, A.N. Samanta, S. Ray, COD reduction of petrochemical industry wastewater using Fentons oxidation, Can. J. Chem. Eng. 88 (2010) 1021 1026. [37] T.J. McGhee, Water Supply and Sewerage, sixth ed., McGraw-Hill, Hightstown, NJ, 1991. [38] H. Salehizadeh, M.C.M. van Loosdrecht, Production of polyhydroxyalkanoates by mixed culture: recent trends and biotechnological importance, Biotechnol. Adv. 22 (2004) 261279. [39] L.A. Fdez-Gelfo, C. lvarez-Gallego, D. Sales, L.I. Romero, The use of thermochemical and biological pretreatments to enhance organic matter hydrolysis and solubilization from organic fraction of municipal solid waste (OFMSW), Chem. Eng. J. 168 (2011) 249254. [40] B. Tang, L. Yu, S. Huang, J. Luo, Y. Zhuo, Energy efciency of pre-treating excess sewage sludge with microwave irradiation, Bioresour. Technol. 101 (2010) 50925097. [41] J. Marin, K.J. Kennedy, C. Eskicioglu, Effect of microwave irradiation on anaerobic degradability of model kitchen waste, Waste Manage. 30 (2010) 17721779. [42] M.L. Torres, M.d.C.E. Llorns, Effect of alkaline pretreatment on anaerobic digestion of solid wastes, Waste Manage. 28 (2008) 22292234. [43] D.C. Devlin, S.R.R. Esteves, R.M. Dinsdale, A.J. Guwy, The effect of acid pretreatment on the anaerobic digestion and dewatering of waste activated sludge, Bioresour. Technol. 102 (2011) 40764082. an, F.D. Sanin, Alkaline solubilization and microwave irradiation as a [44] I. Dog combined sludge disintegration and minimization method, Water Res. 43 (2009) 21392148.

97

[45] J. Kim, C. Park, T. Kim, M. Lee, S. Kim, S. Kim, J. Lee, Effects of various pretreatments for enhanced anaerobic digestion with waste activated sludge, J. Biosci. Bioeng. 95 (2003) 271275. [46] D. Kim, E. Jeong, S. Oh, H. Shin, Combined (alkaline + ultrasonic) pretreatment effect on sewage sludge disintegration, Water Res. 44 (2010) 30933100. [47] A. Cesaro, V. Belgiorno, Sonolysis and ozonation as pretreatment for anaerobic digestion of solid organic waste, Ultrason. Sonochem. 20 (2013) 931936. [48] C. Bougrier, C. Albasi, J.P. Delgens, H. Carrre, Effect of ultrasonic, thermal and ozone pre-treatments on waste activated sludge solubilisation and anaerobic biodegradability, Chem. Eng. Process. 45 (2006) 711718. [49] G. Xu, S. Chen, J. Shi, S. Wang, G. Zhu, Combination treatment of ultrasonic and ozone for improving solubilization and anaerobic biodegradability of waste activated sludge, J. Hazard. Mater. 180 (2010) 340346. [50] C. Eskicioglu, A. Prorot, J. Marin, R.L. Droste, K.J. Kennedy, Synergetic pretreatment of sewage sludge by microwave irradiation in presence of H2O2 for enhanced anaerobic digestion, Water Res. 42 (2008) 46744682. [51] H. Shahriari, M. Warith, M. Hamoda, K.J. Kennedy, Anaerobic digestion of organic fraction of municipal solid waste combining two pretreatment modalities, high temperature microwave and hydrogen peroxide, Waste Manage. 32 (2012) 4152. [52] H.J. Kim, Y.G. Choi, D.Y. Kim, D.H. Kim, T.H. Chung, Effect of pretreatment on acid fermentation of organic solid waste, Water Sci. Technol. 52 (2005) 153 160. [53] S. Bayr, P. Kaparaju, J. Rintala, Screening pretreatment methods to enhance thermophilic anaerobic digestion of pulp and paper mill wastewater treatment secondary sludge, Chem. Eng. J. 223 (2013) 479486. [54] Q. Yang, K. Luo, X. Li, D. Wang, W. Zheng, G. Zeng, J. Liu, Enhanced efciency of biological excess sludge hydrolysis under anaerobic digestion by additional enzymes, Bioresour. Technol. 101 (2010) 29242930. [55] C. Park, C. Lee, S. Kim, Y. Chen, H.A. Chase, Upgrading of anaerobic digestion by incorporating two different hydrolysis processes, J. Biosci. Bioeng. 100 (2005) 164167. [56] L. Appels, S. Houtmeyers, J. Degrve, J.V. Impe, R. Dewil, Inuence of microwave pre-treatment on sludge solubilization and pilot scale semicontinuous anaerobic digestion, Bioresour. Technol. 128 (2013) 598603. [57] J. Ahn, S.G. Shin, S. Hwang, Effect of microwave irradiation on the disintegration and acidogenesis of municipal secondary sludge, Chem. Eng. J. 153 (2009) 145150. [58] M. Climent, I. Ferrer, M.d.M. Baeza, A. Artola, F. Vzquez, X. Font, Effects of thermal and mechanical pretreatments of secondary sludge on biogas production under themophilic conditions, Chem. Eng. J. 133 (2007) 335342. [59] Y. Zhou, M. Takaoka, W. Wang, X. Liu, K. Oshita, Effect of thermal hydrolysis pre-treatment on anaerbic digestion of municipal biowaste: a pilot scale study in China, J. Biosci. Bioeng. 116 (2013) 101105. [60] L. Appels, J. Degrve, B.V.d. Bruggem, J.V. Impe, R. Dewil, Inuence of low temperature thermal pre-treatment on sludge solubilisation heavy metal release and anaerobic digestion, Bioresour. Technol. 101 (2010) 57435748. [61] A. Cesaro, V. Naddeo, V. Amodio, V. Belgiorno, Enhanced biogas production from anaerobic codigestion of solid waste by sonolysis, Ultrason. Sonochem. 19 (2012) 596600. [62] A. Tiehm, K. Nickel, M. Zellhorn, U. Neis, Ultrasonic waste activated sludge disintegration for improving anaerobic stabilization, Water Res. 35 (2001) 20032009. [63] B.R. Dhar, G. Nakhla, M.B. Ray, Techno-economic evaluation of ultrasound and thermal pretreatments for enhanced anaerobic digestion of municipal waste activated sludge, Waste Manage. 32 (2012) 542549. [64] M.R. Salsabil, J. Laurent, M. Casellas, C. Dagot, Techno-economic evaluation of thermal treatment, ozonation and sonication for the reduction of wastewater biomass volume before aerobic or anaerobic digestion, J. Hazard. Mater. 174 (2010) 323333. [65] M. Eddy, Wastewater Engineering: Treatment, disposal and reuse, third ed., Mc-Graw-Hill, Inc., Singapore, 1991. [66] L. Bertin, M.C. Colao, M. Ruzzi, F. Fava, Performances and microbial features of a granular activated carbon packed-bed bioim reactor capable of an efcient anaerobic digestion of olive mill wastewaters, FEMS Microbiol. Ecol. 48 (2004) 413423. [67] C.P.L. Grady, G.T. Daigger, N.G. Love, C.D.M. Filipe, Biological Wastewater Treatment, third ed., CRC Press, Boca Raton, 2011. [68] P.E. Poh, M.F. Chong, Development of anaerobic digestion methods for palm oil mill efuent (POME) treatment, Bioresour. Technol. 100 (2009) 19. [69] Z. Hu, H. Yu, J. Zheng, Application of response surface methodology for optimization of acidogenesis of cattail by rumen cultures, Bioresour. Technol. 97 (2006) 21032109. [70] C. Hong, W. Haiyun, Optimization of volatile fatty acid production with cosubstrate of food wastes and dewatered excess sludge using response surface methodology, Bioresour. Technol. 101 (2010) 54875493. [71] H. Liu, J. Wang, X. Liu, B. Fu, J. Chen, H. Yu, Acidogenic fermentation of proteinaceous sewage sludge: effect of pH, Water Res. 46 (2012) 799807. [72] P.H. Nielsen, A. Jahn, Extraction of EPS, in: J. Wingender, T.R. Neu, H.C. Flemming (Eds.), Microbial Extracellular Polymeric Substances: Characterization, Structure and Function, Springer, Berlin, 1999, p. 58. [73] Y.A. Oktem, O. Ince, T. Donnelly, P. Sallis, B.K. Ince, Determination of optimum operating conditions of an acidication reactor treating a chemical synthesisbased pharmaceutical wastewater, Process Biochem. 41 (2006) 22582263.

98

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 [102] K. Luo, Q. Yang, J. Yu, X. Li, G. Yang, B. Xie, F. Yang, W. Zheng, G. Zeng, Combined effect of sodium dodecyl sulfate and enzyme on waste activated sludge hydrolysis and acidication, Bioresour. Technol. 102 (2011) 7103 7110. [103] H. Liu, J. Wang, A. Wang, J. Chen, Chemical inhibitors of methanogenesis and putative applications, Appl. Microbiol. Biotechnol. 89 (2011) 13331340. [104] S. Philip, T. Keshavarz, I. Roy, Polyhydroxyalkanoates: biodegradable polymers with a range of applications, J. Chem. Technol. Biotechnol. 82 (2007) 233247. [105] J. Choi, S.Y. Lee, Process analysis and economic evaluation for poly(3hydroxybutyrate) production by fermentation, Bioprocess. Biosyst. Eng. 17 (1997) 335342. [106] M.G.E. Albuquerque, M. Eiroa, C. Torres, B.R. Nunes, M.A.M. Reis, Strategies for the development of a side stream process for polyhydroxyalkanoate (PHA) production from sugar cane molasses, J. Biotechnol. 130 (2007) 411421. [107] M.A. Hassan, Y. Shirai, N. Kusubayashi, M.I.A. Karim, K. Nakanishi, K. Hashimoto, The production of polyhydroxyalkanoate from anaerobically treated palm oil mill efuent by Rhodobacter sphaeroides, J. Ferment. Bioeng. 83 (1997) 485488. [108] T. Mumtaz, S. Abd-Aziz, N.A.A. Rahman, P.L. Yee, Y. Shirai, M.A. Hassan, Pilotscale recovery of low molecular weight organic acids from anaerobically treated palm oil mill efuent (POME) with energy integrated system, Afr. J. Biotechnol. 7 (2008) 39003905. [109] A.S.M. Chua, H. Takabatake, H. Satoh, T. Mino, Production of polyhydroxyalkanoates (PHA) by activated sludge treating municipal wastewater: effect of pH, sludge retention time (SRT), and acetate concentration in inuent, Water Res. 37 (2003) 36023611. [110] M.G.E. Albuquerque, C. Torres, M.A.M. Reis, Polyhydroxyalkanoate (PHA) production by a mixed microbial culture using sugar molasses: effect of the inuent substrate concentration on culture selection, Water Res. 44 (2010) 34193433. [111] Y. Jiang, M. Hebly, R. Kleerebezem, G. Muyzer, M.C.M. van Loosdrecht, Metabolic modeling of mixed substrate uptake for polyhydroxyalkanoate (PHA) production, Water Res. 45 (2011) 13091321. [112] S.V. Mohan, M.V. Reddy, Optimization of critical factors to enhance polyhydroxyalkanoates (PHA) synthesis by mixed culture using Taguchi design of experimental methodology, Bioresour. Technol. 128 (2013) 409 416. [113] Y. Jiang, Y. Chen, X. Zheng, Efcient polyhydroxyalkanoates production from a waste-activated sludge alkaline fermentation liquid by activated sludge submitted to the aerobic feeding and discharge process, Environ. Sci. Technol. 43 (2009) 77347741. [114] M.V. Reddy, S.V. Mohan, Inuence of aerobic and anoxic microenvironments on polyhydroxyalkanoates (PHA) production from food waste and acidogenic efuents using aerobic consortia, Bioresour. Technol. 103 (2012) 313321. [115] S. Bengtsson, A.R. Pisco, M.A.M. Reis, P.C. Lemos, Production of polyhydroxyalkanoates from fermented sugar cane molasses by a mixed culture enriched in glycogen accumulating organisms, J. Biotechnol. 145 (2010) 253263. [116] P.C. Lemos, L.S. Seram, M.A.M. Reis, Synthesis of polyhydroxyalkanoates from different short-chain fatty acids by mixed culture submitted to aerobic dynamic feeding, J. Biotechnol. 122 (2006) 226238. [117] P.C. Lemos, C. Viana, N. Salgueiro, A.M. Ramos, J.P.S.G. Crespo, M.A.M. Reis, Effect of carbon source on the formation of polyhydroxyalkanoates (PHA) by a phosphate-accumulating mixed culture, Enzyme Microb. Technol. 22 (1998) 662671. [118] P.A. Holmes, Applications of PHB a microbially produced biodegradable thermoplastic, Phys. Technol. 16 (1985) 3236. [119] Z. Du, H. Li, T. Gu, A state of the art review on microbial fuel cells: a promising technology for wastewater treatment and bioenergy, Biotechnol. Adv. 25 (2007) 464482. [120] G. Mohanakrishna, S.V. Mohan, P.N. Sarma, Utilizing acid-rich efuents of fermentative hydrogen production process as substrate for harnessing bioelectricity: an integrative approach, Int. J. Hydrogen Energy 35 (2010) 34403449. [121] J. Nam, H. Kim, K. Lim, H. Shin, Effects of organic loading rates on the continuous electricity generation from fermented wastewater using a singlechamber microbial fuel cell, Bioresour. Technol. 101 (2010) S33S37. [122] S. Freguia, E.H. Teh, N. Boon, K.M. Leung, J. Keller, K. Rabaey, Microbial fuel cells operating on mixed fatty acids, Bioresour. Technol. 101 (2010) 1233 1238. [123] K. Chae, M. Choi, J. Lee, K. Kim, I.S. Kim, Effect of different substrates on the performance, bacterial diversity, and bacterial viability in microbial fuel cells, Bioresour. Technol. 100 (2009) 35183525. [124] S. Teng, Z. Tong, W. Li, S. Wang, G. Sheng, X. Shi, X. Liu, H. Yu, Electricity generation from mixed volatile fatty acids using microbial fuel cells, Appl. Microbiol. Biotechnol. 87 (2010) 23652372. [125] W. Lv, F.L. Schanbacher, Z. Yu, Putting microbes to work in sequence: recent advances in temperature-phased anaerobic digestion processes, Bioresour. Technol. 101 (2010) 94099414. [126] G.N. Demirer, S. Chen, Two-phase anaerobic digestion of unscreened dairy manure, Process Biochem. 40 (2005) 35423549. [127] C. Cavinato, D. Bolzonella, F. Fatone, F. Cecchi, P. Pavan, Optimization of twophase thermophilic anaerobic digestion of biowaste for hydrogen and

[74] J.-I. Horiuchi, T. Shimizu, K. Tada, T. Kanno, M. Kobayashi, Selective production of organic acids in anaerobic acid reactor by pH control, Bioresour. Technol. 82 (2002) 209213. [75] J. Horiuchi, T. Shimizu, T. Kanno, M. Kobayashi, Dynamic behavior in response to pH shift during anaerobic acidogenesis with a chemostat culture, Biotechnol. Tech. 13 (1999) 155157. [76] H.Q. Yu, H.H.P. Fang, Acidogenesis of gelatin-rich wastewater in an upow anaerobic reactor: inuence of pH and temperature, Water Res. 37 (2003) 55 66. [77] H.-Q. Yu, H.H.P. Fang, Acidogenesis of dairy wastewater at various pH levels, Water Sci. Technol. 45 (2002) 201206. [78] Q. Yuan, R. Sparling, J.A. Oleszkiewicz, VFA generation from waste activated sludge: effect of temperature and mixing, Chemosphere 82 (2011) 603607. [79] P. Zhang, Y. Chen, T. Huang, Q. Zhou, Waste activated sludge hydrolysis and short-chain fatty acids accumulation in the presence of SDBD in semicontinuous ow reactors: effect of solids retention time and temperature, Chem. Eng. J. 148 (2009) 348353. [80] G. Zhuo, Y. Yan, X. Tan, X. Dai, Q. Zhou, Ultrasonic-pretreated waste activated sludge hydrolysis and volatile fatty acid accumulation under alkaline conditions: effect of temperature, J. Biotechnol. 159 (2012) 2731. [81] A. Bouzas, C. Gabaldn, P. Marzal, J.M. Penya-roja, A. Seco, Fermentation of municipal primary sludge: effect of SRT and solids concentration on volatile fatty acid production, Environ. Technol. 23 (2002) 863875. [82] H. Yu, H.H.P. Fang, G. Gu, Comparative performance of mesophilic and thermophilic acidogenic upow reactors, Process Biochem. 38 (2002) 447 454. [83] J. Yu, M. Zheng, T. Tao, J. Zuo, K. Wang, Waste activated sludge treatment based on temperature staged and biologically phased (TSBP) anaerobic digestion system, J. Environ. Sci., http://dx.doi.org/10.1016/S10010742(12)60266-6. [84] J. Lu, B.K. Ahring, Effects of temperature and hydraulic retention time on thermophilic anaerobic pretreatment of sewage sludge, in: 4th International Symposium on Anaerobic Digestion of Solid Waste, Copenhagen, 2005, pp. 159164. [85] D. Bolzonella, P. Pavan, M. Zanette, F. Cecchi, Two-phase anaerobic digestion of waste activated sludge: effect of an extreme thermophilic prefermentation, Ind. Eng. Chem. Res. 46 (2007) 66506655. [86] J. Lu, H.N. Gavala, I.V. Skiadas, Z. Mladenovska, B.K. Ahring, Improving anaerobic sewage sludge digestion by implementation of a hyperthermophilic prehydrolysis step, J. Environ. Manage. 88 (2008) 881889. [87] S.I. Gadow, H. Jiang, R. Watanabe, Y. Li, Effect of temperature and temperature shock on the stability of continuous cellulosic-hydrogen fermentation, Bioresour. Technol. 142 (2013) 304311. [88] M. Ben, T. Mato, A. Lopez, M. Vila, C. Kennes, M.C. Veiga, Bioplastic production using wood mill efuents as feedstock, Water Sci. Technol. 63 (2011) 1196 1202. [89] S. Lim, B.J. Kim, C. Jeong, J. Choi, Y.H. Ahn, H.N. Chang, Anaerobic organic acid production of food waste in once-a-day feeding and drawing-off bioreactor, Bioresour. Technol. 99 (2008) 78667874. [90] H.H.P. Fang, H.Q. Yu, Effect of HRT on mesophilic acidogenesis of dairy wastewater, J. Environ. Eng. 126 (2000) 11451148. [91] R.M. Dinsdale, G.C. Premier, F.R. Hawkes, D.L. Hawkes, Two-stage codigestion of waste activated sludge and fruit/vegetable waste using inclined tubular digesters, Bioresour. Technol. 72 (2000) 159168. [92] H. Xiong, J. Chen, H. Wang, H. Shi, Inuences of volatile solid concentration, temperature and solid retention time for the hydrolysis of waste activated sludge to recover volatile fatty acids, Bioresour. Technol. 119 (2012) 285292. [93] L. Feng, H. Wang, Y. Chen, Q. Wang, Effect of solids retention time and temperature on waste activated sludge hydrolysis and short-chain fatty acids accumulation under alkaline conditions in continuous-ow reactors, Bioresour. Technol. 100 (2009) 4449. [94] Y. Miron, G. Zeeman, J.B.V. Lier, G. Lettinga, The role of sludge retention time in the hydrolysis and acidication of lipids, carbohydrates and proteins during digestion of primary sludge in CSTR systems, Water Res. 34 (2000) 17051713. [95] I. Ferrer, F. Vzquez, X. Font, Long term operation of a thermophilic anaerobic reactor: process stability and efciency at decreasing sludge retention time, Bioresour. Technol. 101 (2010) 29722980. [96] Z. Salmiati, M.R. Ujang, M.F.M. Salim, M.A. Din, Ahmad, Intracellular biopolymer productions using mixed microbial cultures from fermented POME, Water Sci. Technol. 56 (2007) 179185. [97] Q. Yuan, R. Sparling, J.A. Oleszkiewicz, Waste activated sludge fermentation: effect of solids retention time and biomass concentration, Water Res. 43 (2009) 51805186. [98] J. Yu, Production of PHA from starchy wastewater via organic acids, J. Biotechnol. 86 (2001) 105112. [99] B. Rincn, E. Snchez, F. Raposo, R. Borja, L. Travieso, M.A. Martn, A. Martn, Effect of organic loading rate on the performance of anaerobic acidogenic fermentation of two-phase olive mill solid residue, Waste Manage. 28 (2008) 870877. [100] E. Nebot, L.I. Romero, J.M. Quiroga, D. Sales, Effect of the feed frequency on the performance of anaerobic lters, Anaerobe 1 (1995) 113120. [101] H. Liu, H.H.P. Fang, Extraction of extracellular polymeric substances (EPS) of sludges, J. Biotechnol. 95 (2002) 249256.

W.S. Lee et al. / Chemical Engineering Journal 235 (2014) 8399 methane production through reject water recirculation, Bioresour. Technol. 102 (2011) 86058611. W. Wang, L. Xie, J. Chen, G. Luo, Q. Zhou, Biohydrogen and methane production by co-digestion of cassava stillage and excess sludge under thermophilic condition, Bioresour. Technol. 102 (2011) 38333839. T. Dogan, O. Ince, N.A. Oz, B.K. Ince, Inhibition of volatile fatty acid production in granular sludge from a UASB reactor, J. Environ. Sci. Health, Part A: Toxic/ Hazard. Subst. Environ. Eng. 40 (2005) 633644. Y. Wang, Y. Zhang, J. Wang, L. Meng, Effects of volatile fatty acid concentrations on methane yield and methanogenic bacteria, Biomass Bioenergy 33 (2009) 848853. W. Zong, R. Yu, P. Zhang, M. Fan, Z. Zhou, Efcient hydrogen gas production from cassava and food waste by a two-step process of dark fermentation and photo-fermentation, Biomass Bioenergy 33 (2009) 14581463. E. Tuna, F. Kargi, H. Argun, Hydrogen gas production by electrohydrolysis of volatile fatty acid (VFA) containing dark fermentation efuent, Int. J. Hydrogen Energy 34 (2009) 262269. W. Liu, S. Huang, A. Zhou, G. Zhou, N. Ren, A. Wang, G. Zhuang, Hydrogen generation in microbial electrolysis cell feeding with fermentation liquid of waste activated sludge, Int. J. Hydrogen Energy 37 (2012) 1385913864. D.B. Levin, L. Pitt, M. Love, Biohydrogen production: prospects and limitations to practical application, Int. J. Hydrogen Energy 29 (2004) 173185. C. Chen, M. Yang, K. Yeh, C. Liu, J. Chang, Biohydrogen production using sequential two-stage dark and photo fermentation processes, Int. J. Hydrogen Energy 33 (2008) 47554762. H. Su, J. Cheng, J. Zhou, W. Song, K. Cen, Combination of dark- and photofermentation to enhance hydrogen production and energy conversion efciency, Int. J. Hydrogen Energy 34 (2009) 88468853. E. zgr, A.E. Mars, B. Peksel, A. Louwerse, M. Ycel, U. Gndz, P.A.M. _ Erog lu, Biohydrogen production from beet molasses by sequential Claassen, I. dark and photofermentation, Int. J. Hydrogen Energy 35 (2010) 511517. A.J. Guwy, R.M. Dinsdale, J.R. Kim, J. Massanet-Nicolau, G. Premier, Fermentative biohydrogen production systems integration, Bioresour. Technol. 102 (2011) 85348542. C. Chen, K. Yeh, Y. Lo, H. Wang, J. Chang, Engineering strategies for the enhanced photo-H2 production using efuents of dark fermentation processes as substrate, Int. J. Hydrogen Energy 35 (2010) 1335613364. K. Vijayaraghavan, G.K. Sagar, Anaerobic digestion and in situ electrohydrolysis of dairy bio-sludge, Biotechnol. Bioprocess Eng. 15 (2010) 520526. H. Liu, S. Grot, B.E. Logan, Electrochemically assisted microbial production of hydrogen from acetate, Environ. Sci. Technol. 39 (2005) 43174320. F. Xue, J. Miao, X. Zhang, H. Luo, T. Tan, Studies on lipid production by Rhodotorula glutinis fermentation using monosodium glutamate wastewater as culture medium, Bioresour. Technol. 99 (2008) 59235927. M.M. Gui, K.T. Lee, S. Bhatia, Feasibility of edible oil vs. non-edible oil vs. waste edible oil as biodiesel feedstock, Energy 33 (2008) 16461653. Z. Chi, Y. Zheng, J. Ma, S. Chen, Oleaginous yeast Cryptococcus curvatus culture with dark fermentation hydrogen production efuent as feedstock for microbial lipid production, Int. J. Hydrogen Energy 36 (2011) 95429550. Q. Fei, H.N. Chang, L. Shang, J. Choi, N. Kim, J. Kang, The effect of volatile fatty acids as a sole carbon source on lipid accumulation by Cryptococcus albidus for biodiesel production, Bioresour. Technol. 102 (2011) 26952701. Q. Fei, H.N. Chang, L. Shang, J. Choi, Exploring low-cost carbon sources for microbial lipids production by fed-batch cultivation of Cryptococcus albidus, Biotechnol. Bioprocess Eng. 16 (2011) 482487.

99

[128]

[129]

[130]

[131]

[132]

[133]

[134] [135]

[136]

[137]

[138]

[139]

[140]

[141] [142]

[143] [144]

[145]

[146]

[147] M. Henze, Capabilities of biological nitrogen removal processes from wastewater, Water Sci. Technol. 23 (1991) 669679. [148] J. Tong, Y. Chen, Enhanced biological phosphorus removal driven by shortchain fatty acids produced from waste activated sludge alkaline fermentation, Environ. Sci. Technol. 41 (2007) 71267130. [149] S.-J. Lim, D.W. Choi, W.G. Lee, S. Kwon, H.N. Chang, Volatile fatty acids production from food wastes and its application to biological nutrient removal, Bioprocess. Biosyst. Eng. 22 (2000) 543545. [150] P. Elefsiniotis, D.G. Wareham, Utilization patterns of volatile fatty acids in the denitrication reaction, Enzyme Microb. Technol. 41 (2007) 9297. [151] Y. Chen, A.A. Randall, T. McCue, The efciency of enhanced biological phosphorus removal from real wastewater affected by different ratios of acetic to propionic acid, Water Res. 38 (2004) 2736. [152] L.T. Angenent, K. Karim, M.H. Al-Dahhan, B.A. Wrenn, R. Domguez-Espinosa, Production of bioenergy and biochemicals from industrial and agricultural wastewater, Trends Biotechnol. 22 (2004) 477485. [153] B.E. Logan, Microbial Fuel Cells, John Wiley & Sons, Inc., Hoboken, New Jersey, 2008. [154] T. Mato, M. Ben, C. Kennes, M.C. Veiga, Valuable product production from wood mill efuents, Water Sci. Technol. 62 (2010) 22942300. [155] A.S. Ucisik, M. Henze, Biological hydrolysis and acidication of sludge under anaerobic conditions: the effect of sludge type and origin on the production and composition of volatile fatty acids, Water Res. 42 (2008) 3729 3738. [156] A. Banerjee, P. Elefsiniotis, D. Tuhtar, The effect of addition of potatoprocessing wastewater on the acidogenesis of primary sludge under varied hydraulic retention time and temperature, J. Biotechnol. 72 (1999) 203 212. [157] L. Feng, Y. Yan, Y. Chen, Co-fermentation of waste activated sludge with food waste for short-chain fatty acids production: effect of pH at ambient temperature, Front. Environ. Sci. Eng. China 5 (2011) 623632. [158] K.S. Min, A.R. Khan, M.K. Kwon, Y.J. Jung, Y. Kiso, Acidogenic fermentation of blended food-waste in combination with primary sludge for the production of volatile fatty acids, J. Chem. Technol. Biotechnol. 80 (2005) 909915. [159] E. Alkaya, G.N. Demirer, Anaerobic acidication of sugar-beet processing wastes: effect of operational parameters, Biomass Bioenergy 35 (2011) 32 39. [160] H. Yu, Z. Wang, Q. Wang, Z. Wu, J. Ma, Disintegration and acidication of MBR sludge under alkaline conditions, Chem. Eng. J. 231 (2013) 206213. [161] T. Mumtaz, N.A. Yahaya, S. Abd-Aziz, N.A.A. Rahman, P.L. Yee, Y. Shirai, M.A. Hassan, Turning waste to wealth-biodegradable plastics polyhydroxyalkanoates from palm oil mill efuent a Malaysian perspective, J. Clean Prod. 18 (2010) 13931402. [162] M. Zhang, H. Wu, H. Chen, Coupling of polyhydroxyalkanoate production with volatile fatty acid from food wastes and excess sludge, Process Saf. Environ. Protect., http://dx.doi.org/10.1016/j.psep.2012.12.002. [163] P. Cavdar, E. Yilmaz, A.E. Tugtas, B. Calli, Acidogenic fermentation of municipal solid waste and its application to bio-electricity production via microbial fuel cells (MFCs), Water Sci. Technol. 64 (2011) 789 795. [164] J. Tong, Y. Chen, Recovery of nitrogen and phosphorus from alkaline fermentation liquid of waste activated sludge and application of the fermentation liquid to promote biological municipal wastewater treatment, Water Res. 43 (2009) 29692976.

You might also like