Acessibilidade / Reportar erro

Comparison of acid and enzymatic hydrolysis of pectin, as inexpensive source to cell growth of Cupriavidus necator

Abstract

Abstract: The present work investigated what the appropriate methods of hydrolysis of pectin for reducing compounds (RCs) production, employed as a substrate for cell growth of Cupriavidus necator. This microorganism has great importance industrial, because besides potential single cell protein (SCP), is the most studied microorganism for production of polyhydroxybutyrate (PHB), and both processes require high cell concentration with inexpensive substrates For this, it was compared to acid and enzymatic hydrolysis procedures, through rotational central composite experimental design, using pectin concentration (1.0%). It was analyzed as a variable response for both experimental design, the RCs’ production. The best conditions of each procedure were used in study kinetics of RCs’ production and as a substrate for cell growth of C. necator. The results indicated that the enzymatic hydrolysis method was the most efficient, with a 93.0% yield of RCs, while the yield for acid hydrolysis was 60.0%. The optimum conditions for enzymatic hydrolysis were an enzyme concentration of 10.01 UI/g (International Unit of enzyme per gram of pectin) and an agitation speed of 230.3 rpm. C. necator showed satisfactory growth in the media containing pectin hydrolysates, with specific growth rates (μMax) similar to those reported for other substrates.

Key words
Cupriavidus necator; galacturonic acid; pectin depolymerization; pectin hydrolysates; polygalacturonase


INTRODUCTION

Pectins are among the most abundant natural polysaccharides, present as a component of the primary cell wall and middle lamella of fruits and vegetables and normally found associated with cellulose, hemicellulose, and lignin. Although ubiquitous in almost all plants, pectins of the citrus, apple, sugar beet, and sunflower are considered of special interest, due to the physicochemical quality and the availability of their biomasses in agroindustrial wastes (MuzzarelliMUZZARELLI RAA, BOUDRANT J, MEYER D, MANNO N, DEMARCHIS M and PAOLETTI MG. 2012. Current views on fungal chitin/chitosan, human chitinases, food preservation, glucans, pectins and inulin: A tribute to Henri Braconnot, precursor of the carbohydrate polymers science, on the chitin bicentennial. Carbohydr Polym 87: 995-1012. et al. 2012, AdetunjiADETUNJI LR, ADEKUNLE A, ORSAT V and RAGHAVAN V. 2017. Advances in the pectin production process using novel extraction techniques: A review. Food Hydrocoll 62: 239-250. et al. 2017).

Pectins are high molecular weight heteropolymers, with a high content of galacturonic acid (GalA) – an oxidized form of D-galactose, which constitutes the main monomeric unit (around 65.0%) of the pectin molecule. The structure of pectins can change according to the material of origin, and its understanding is currently characterized by much speculation and has not yet been fully resolved. A general model proposes linearly alternating chains of homogalacturonans, composed of repeating units of (1→4)-α-D-GalA, interrupted by branched regions composed of (1→2)-α-L-rhamnose units, to which are bound neutral sugars including galactose, arabinose, xylose, and fructose (MayMAY CD. 2000. Pectins. In: Phillips GO and Williams PA (Eds), Handbook of Hydrocolloids, New York: CRC Press, p. 165-185. 2000, CaffallCAFFALL KH and MOHNEN D. 2009. The structure, function, and biosynthesis of plant cell wall pectic polysaccharides. Carbohydr Res 344: 1879-1900. and Mohnen 2009).

Pectin can be degraded by either acid or enzymatic hydrolysis. Acid methods are commonly used in analytical procedures; however, divergent results suggest that these techniques still require significant improvement. The maintenance of strongly acidic conditions for prolonged periods can result in a rate of destruction of free GalA that exceeds the rate of polymer release (GarnaGARNA H, MABON N, NOTT K, WATHELET B and PAQUOT M. 2006. Kinetic of the hydrolysis of pectin galacturonic acid chains and quantification by ionic chromatography. Food Chem 96: 477-484. et al. 2006).

Enzymatic procedures are widely used in industry to improve the yield during the extraction and clarification of juices. The (endo)polygalacturonases (E.C.3.2.1.15) are probably the most important pectinases for biocatalysis, since they are able to hydrolyze both pectin and pectic acids. However, few kinetic studies have been reported concerning the mode of action of these enzymes (KissKISS K, CSERJÉSI P, NEMESTÓTHY N, GUBICZA L and BÉLAFI-BAKÓ K. 2008. Kinetic study on hydrolysis of various pectins by Aspergillus niger polygalacturonase. Hung J Ind Chem 36: 55-58. et al. 2008, KohliKOHLI P and GUPTA R. 2015. Alkaline pectinases: A review. Biocatal Agric Biotechnol 4: 279-285. and Gupta 2015).

Agroindustrial pectin-rich waste, such as fruit pulp, husks, and bagasse, are potentially suitable feedstocks for bioconversion into products of biotechnological interest. Often bioconversions of these residues can be done by acid or enzymatic hydrolysis to provide a useful source of carbon and energy for use in biotransformation processes (PintoPINTO GAS, BRITO ES, SILVA FLH, SANTOS SFM and MACEDO GR. 2006. Fermentação em estado sólido: Uma alternativa para o aproveitamento e valorização de resíduos agroindustriais. Rev Quim Ind 74: 17-20. et al. 2006).

Although few studies have been undertaken concerning the saccharification of pectin, D-galacturonic acid is an important primary material in the food, pharmaceutical, and cosmetic industries and can be used to produce vitaminMIN B, LIM J, KO S, LEE KG, LEE SH and LEE S. 2011. Environmentally friendly preparation of pectins from agricultural byproducts and their structural / rheological characterization. Bioresour Technol 102: 3855-3860. C, acidification agents, and surfactants (JörnedingJÖRNEDING HJ, BACIO IE, BERENSMEYER S and BUCHHOLZ K. 2002. Gewinnung von galacturonsäure aus zellwandbestandteilen der rübenschnitzel. Zuckerind 127: 845-853. et al. 2002). As an application alternative, previous work by our research group demonstrated the capability of C. necator to grow using GalA as the sole source of carbon as well as using the products of acid hydrolysis of pectin (LocatelliLOCATELLI GO, SILVA GD, FINKLER L and FINKLER CLL. 2011. Acid hydrolysis of pectin for cell growth of Cupriavidus necator. Biotechnol 1: 1-8. et al. 2011), but optimizing the hydrolysis conditions should be further studied.

C. necator was famous as a potential single cell protein (SCP) in the 1970s, studies evaluated that bacterial cells would average 50.0% protein, with 93.0% of digestibility by animals and high concentration of important amino acids similar to those found in casein. However, the competition from soybased protein resulted in SCP not receiving much attention, but in recent years has resurgence the interest in SCP and PHA as a component of animal feed to increase the metabolizable energy content (KunasundariKUNASUNDARI B, MURUGAIYAH V, KAUR G, MAUER FHJ and SUDESH K. 2013. Revisiting the Single Cell Protein Application of Cupriavidus necator H16 and Recovering Bioplastic Granules Simultaneously. Plos One 8: 1-15. et al. 2013).

Although hundreds of species of microorganisms are capable of accumulating polyhydroxyalkanoates (PHAs), C. necator is the most important microorganism and has been extensively studied for industrial production of PHAs because it can accumulate up to 80.0% of its dry mass in the form of biopolymers (AkaraonyeAKARAONYE E, KESHAVARZ T and ROY I. 2010. Production of polyhydroxyalkanoates: the future green materials of choice. J Chem Technol Biotechnol 85: 732-743. et al. 2010, WangWANG Y, YIN J and CHEN GQ. 2014. Polyhydroxyalkanoates, challenges and opportunities. Curr Opin Biotechnol 30: 59-65. et al. 2014). However, the polyhydroxyalkanoates production is still 5 to 10 times more expensive than chemically synthesized polymers, and the substrate may represent more than 40.0% of the cost of production (Akaraonye et al. 2010, AlbuquerqueALBUQUERQUE PBS and MALAFAIA CB. 2018. Perspectives on the production, structural characteristics and potential applications of bioplastics derived from polyhydroxyalkanoates. Int J Biol Macromol 107: 615-625. and Malafaia 2018).

Thus, either for SCP production or for PHA production is very important to identify a new and inexpensive substrates source that can be used to cell growth of C. necator. Like this, the objective this work was to compare methods of acid and enzymatic hydrolysis for the saccharification of pectin and investigate its use as a substrate for the growth of C. necator.

MATERIALS AND METHODS

A commercial citrus pectin with a high degree of esterification (67.0%) was purchased (Vetec, Brazil). Sulfuric acid (Vetec, Brazil) was employed in the acid hydrolysis and a polygalacturonase (Sigma-Aldrich, St. Louis, US, E.C.3.2.1.15), with an enzymatic activity of 1.32 UI per mg of enzyme, was used in the enzymatic hydrolysis, the value of enzymatic activity was confirmed experimentally (data not shown).

In order to refine the conditions for pectin hydrolysis and maximize the yield of reducing compounds (RCs), two independent variables (defined separately for each hydrolysis method) were evaluated using a full 22 experimental design (rotational central composite design – RCCD), with three central points (level 0) and four axial points (levels ± α, where α = 1.4142), totaling 11 experiments. This experimental design model, it allows a greater comprehension of the parameters tested, minimizing the experiments number. The experiments were performed randomly, and the data were analyzed using Statistica 8.0 software (StatSoft, Dell Software, US), with a 95.0% confidence level. The experimental error was obtained from the mean and standard deviation of the central points. The software calculate an empirical model described by Equation, through the experimental data, allowing prediction of the experimental value at any point within the study area.

ACID HYDROLYSIS

The acid hydrolysis of pectin was based on the method proposed by WenzelWENZEL GE. 2001. Bioquímica Experimental dos Alimentos, 1a ed., Unisinos, Rio Grande do Sul, 213 p. (2001). The tests were performed using a rotary evaporator reflux system (Marconi). The initial pectin concentration was 1.0% (w/v), and temperature and acid concentration were the independent variables (Table I). For this, a 5.0 g portion of pectin was added to 500 mL of a solution of sulfuric acid in a distillation flask, and the mixture was refluxed for 5 hours. At the end of each test, the hydrolysate was cooled in an ice bath and immediately neutralized with NaOH 50.0% (w/v).

TABLE I
Codified levels and actual values of the variables studied in the acid hydrolysis experiments.

The results of the experimental design were used to define the hydrolysis conditions that favored RCs’ production. Experiments were then performed in triplicate under these conditions, with samples removed at the start, and then after every 15 minutes during the first hour, and subsequently after every 30 minutes up to a final time of 5 hours. The samples were immediately neutralized with NaOH 50.0% (w/v), cooled in an ice bath, diluted for analysis of RCs, and stored at -18 oC until the chromatographic procedures were performed. The final hydrolysate obtained was neutralized, sterilized in an autoclave, and used as the substrate in the microorganism culture medium.

ENZYMATIC HYDROLYSIS

According to the information provided by Sigma, the polygalacturonase utilized for the enzymatic hydrolysis experiments presented optimum activity at pH 4.0 and a temperature of 50 oC. To enable comparison with the acid hydrolysis method, the same pectin concentration (1.0% w/v) was employed, and the independent variables were the enzyme concentration and the agitation speed (Table II).

TABLE II
Codified levels and actual values of the variables studied in the enzymatic hydrolysis experiments.

The tests were performed using 125 mL Erlenmeyer flasks containing 0.5 g of pectin and 50 mL of sodium acetate buffer (50 mM) and incubated under orbital agitation for 24 hours. At the end of the reaction period, the enzyme activity was interrupted by placing the sample in a boiling water bath for 5 minutes.

The optimum conditions for enzymatic hydrolysis were identified from the results, and experiments were then performed in triplicate, with removal of sample aliquots at the start of the experiment, and then after every 15 minutes during the first hour, and subsequently after every hour during a total period of 24 hours. After removal, the samples were immediately placed in a boiling water bath for 5 minutes, diluted for analysis of RCs, and stored at -18 oC until the chromatographic procedures were performed. The final hydrolysate was neutralized, sterilized in an autoclave, and it was used as the substrate in the microorganism culture medium.

MICROORGANISM AND MAINTENANCE OF THE CULTURE

The strain of Cupriavidus necator was obtained from the Deutsche Sammlung von Mikroorganismen und Zellkulturen (DSMZ 545), and maintained in the culture collection of the Department of Antibiotics of the Federal University of Pernambuco (UFPEDA 0604). During the tests, the bacterium culture was frequently subcultured on tubes containing nutrient agar slants and stored in a refrigerator (4–8 oC).

PREPARATION OF THE INOCULUM

For the production of the inoculum, one loopful of the bacterial culture was transferred from a slant culture into an Erlenmeyer flask (250 mL) containing 100 mL nutrient broth (NB) medium. The flask was incubated in a shaking incubator at 30 °C and 300 rpm for 10 hours. This time of cultivation had previously been established for attainment of the exponential growth phase (data not shown).

CULTURE CONDITIONS

The culture was performed under the same conditions described previously, employing an inoculum of 5.0% (v/v) of the cellular material obtained in the previous step. The mineral medium used was described by RamsayRAMSAY BA, LOMALIZA K, CHAVARIE C, DUBE B, BATAILLE P and RAMSAY JA. 1990. Production of poly-(β-hydroxybutyric-co-β-hydroxyvaleric) acids. Appl Environ Microbiol 56: 2093-2098. et al. (1990), modified by AragãoARAGÃO GMF, LINDLEY ND, URIBELARREA JL and PAREILLEUX A. 1996. Maintaining a controlled residual growth capacity increases the production of polyhydroxyalkanoate copolymers by A. eutrophus. Biotechnol Lett 18: 937-942. et al. (1996). The medium was composed of a mixture of four solutions, as follows (with concentrations as g/L): Nitrolactic acid (0.19), ferrous ammonium citrate (0.06), MgSO4.7H2O (0.5), CaCl2.2H2O (0.01), solution of oligoelements (1.0 mL) (Solution 1); Na2HPO4.12H2O (8.95), KH2PO4 (1.5) (Solution 2); (NH4)2SO4 (5.0) (Solution 3); pectin hydrolysate (Solution 4). The solution of oligoelements consisted of (g/L): H3BO3 (3.0), CoCl2.6H2O (0.2), ZnSO4.7H2O (0.1), MnCl2.4H2O (0.03), Na2MoO4.2H2O (0.03), NiCl2.6H2O (0.02), and CuSO4.5H2O (0.01).

The pH of the solutions was adjusted to 7.0 with KOH (5.0 M). The solutions were then autoclaved separately and mixed aseptically to produce the medium. Samples were withdrawn at intervals of 2 hours for analyses of pH, cell concentration, and consumption of RCs.

ANALYTICAL METHODS

Cell growth was measured using a Marconi spectrophotometer, operated at a wavelength of 600 nm. The optical density values were correlated to the dry mass using a calibration curve. The determination of bacterial cell dry mass was performed by drying in an oven at 70 oC to a constant weight, after filtration using a 0.22 μm membrane. The pH was monitored using a Marconi potentiometer. Measurements of RCs were performed according to the 3,5-dinitrosalicylic acid (DNS) method described by MillerMILLER GL. 1959. Use of dinitrosalicylic acid reagent for determination of reducing sugar. Anal Chem 31: 426-428. (1959). Glucose was used as a standard to produce the calibration curve.

The RCs were identified by high-performance liquid chromatography (Varian, Walnut Creek, CA, US), using a refractive index detector. A column suitable for organic acids was employed (Aminex HPX – 87H, 300 x 780 mm, Bio-Rad, Hercules, CA, US), maintained at 65 ºC (KleinKLEIN H and LEUBOLT R. 1993. Ion-exchange high-performance liquid chromatography in the brewing industry. J Chromatogr 640: 259-270. and Leubolt, 1993). The mobile phase was an aqueous solution of H2SO4 (8.0 mM), and the flow rate was 0.6 mL/min. Retention times were determined using standard solutions of GalA, fructose, galactose, xylose, rhamnose and arabinose (Sigma), at concentrations in the range 0.2–10 g/L. The RCs’ concentrations were expressed as a separate galacturonic acid group (GalA) and the sum of neutral sugars (NeutralS), compounds for fructose + xylose + galactose + rhamnose + arabinose.

RESULTS AND DISCUSSION

ACID HYDROLYSIS OPTIMIZATION

RCs’ concentrations for different experiment conditionals were evaluated based on results of experimental design. The best experimental results obtained was 4.6 g/L of RCs, with 1.0% (v/v) sulfuric acid at 85 oC, but the empirical model predict increase RCs’ production with 1.0% (v/v) sulfuric acid at 100 oC (Figure 1a). The Pareto chart (Figure 1c) and ANOVA (Table III) showed that all of the effects were significant (p < 0.05), and were therefore used in the prediction of an empirical model described by Equation 1. The parameter (1) L – Concentration of H2SO4 (% v/v) exerted the greatest influence on RC production, but this effect was negative, decreasing RC production with increasing of concentration of H2SO4. On the other hand, the parameters (2) L – Temperature (oC) and Q – Temperature (oC) exerted a positive influence on RC production. The distribution of the residuals (values predicted by the model versus observed values) showed that the deviations were normally distributed, and that there was a satisfactory correlation between the theoretical and experimental values (Figure 1b).

R C = 24.892 + 1.705 E n z 0.067 E n z 2 + 0.219 A g i t 0.00044 E n z 2 < i t a l i c > R C = 14.870 + 0.169 H 2 S O 4 0.019 ( H 2 S O 4 ) 2 0.292 T + 0.002 T 2 0.005 H 2 S O 4 T (1)
Figure 1
Experimental design to pectin acid hydrolysis optimization – response surface (a), distribution of residuals (b), and Pareto chart (c) as a function of independent variables: temperature and H2SO4 concentration, and response variable: reducing compounds (RCs) production.
TABLE III
Acid hydrolysis optimization - ANOVA statistics data of independent variables and their interactions; dependent variable (RC concentration g/L).

Therefore, the formation of RCs was favored by decreasing the sulfuric acid concentration and increasing the temperature. LeitãoLEITÃO MCA, ALARCÃO-SILVA ML, JANUÁRIO MIN and AZINHEIRA HG. 1995. Galacturonic acid in pectic substances of sunflower head residues: quantitative determination by HPLC. Carbohydr Polym 26: 165-169. et al. (1995), who used hydrochloric acid and trifluoroacetic acid to hydrolyze sunflower pectin, obtained similar results. These authors observed that the GalA yield increased at higher temperatures and lower acid concentrations. Garna et al. (2006) studied the hydrolysis of pectin using different concentrations of sulfuric acid and achieved the best results at acid concentrations around 1.0 M at 100 oC and lower hydrolysis rates using acid concentrations of 0.2 and 2.0 M. Other researchers also indicated the high stability of PGalA in acidic environments (LimLIM J, YOO J, KO S and LEE S. 2012. Extraction and characterization of pectin from Yuza (Citrus junos) pomace: A comparison of conventional-chemical and combined physical-enzymatic extractions. Food Hydrocoll 29: 160-165., et al. 2012, Min et al. 2011).

Garna et al. (2006) also observed the positive effect of temperature on the yield of free GalA. The positive influence of temperature on hydrolysis can be explained by greater solubilization of the polysaccharideDe RUITER GA, SCHOLS HA, VORAGEN AGJ and ROMBOUTS FM. 1992. Carbohydrate analysis of water-soluble uronic acid containing polysaccharides with high-performance anion-exchange chromatography using methanolysis combined with TFA hydrolysis is superior to four other methods. Anal Biochem 207: 176-185.s at higher temperatures, and the greater resistance of the glycosidic bonds under milder hydrolysis conditions (BiermannBIERMANN CJ. 1988. Hydrolysis and other cleavages of glycosidic linkages in polysaccharides. Adv Carbohydr Chem Biochem 46: 251-272. 1988, De Ruiter et al. 1992). The negative effect of high acid concentrations can be explained by the decomposition of the RCs by other products. The combination of acid and high temperature may cause the formation of furfural derivatives, resulting in an imprecise determination of sugars (WikieraWIKIERA A, MIKA M, STARZYNSKA-JANISZEWSKA A and STODOLAK B. 2015. Development of complete hydrolysis of pectins from apple pomace. Food Chem 172: 675-680. et al. 2015). Such secondary reactions not only reduce the yield of the desired monosaccharides but also produce toxic compounds that prohibit the use of these hydrolysates in biological conversion processes.

The results clearly showed that optimum hydrolysis was achieved at higher temperatures and acid concentrations of up to 1.0% (v/v). Since it was not practically feasible to raise the temperature above 100 oC, the conditions chosen for the RC-release kinetics experiments were a temperature of 100 oC and an acid concentration of 1.0% (v/v) H2SO4.

ENZYMATIC HYDROLYSIS OPTIMIZATION

The best experimental results obtained was 8.8 g/L of RCs, with 11 UI/g of pectin and an agitation speed of 211.6 rpm, but the empirical model predict maximum concentration of RCs can be achieved using an enzyme concentration of 10.01 UI/g of pectin and an agitation speed of 230.3 rpm (Figure 2a). All of the effects were significant (p < 0.05) as showed the Pareto chart (Figure 2c) and ANOVA (Table IV), the observation that the second order effects were negative indicated that there was an optimum point for the two variables. Figure 2b shows that there was a good agreement between experimental data and numerical predictions. The codified model, optimized for the enzymatic hydrolysis of pectin after 24 hours of bioreaction, is given by Equation 2.

R C = 24.892 + 1.705 E n z 0.067 E n z 2 + 0.219 A g i t 0.00044 E n z 2 (2)
Figure 2
Experimental design to pectin enzymatic hydrolysis optimization - response surface (a), distribution of residuals (b), and Pareto chart (c) as a function of independent variables: agitation speed and enzyme concentration, and response variable: reducing compounds (RCs) production.
TABLE IV
Enzymatic hydrolysis optimization - ANOVA statistics data of independent variables and their interactions; dependent variable (RC concentration g/L).

Thus, can conclude that both variables have a great influence on the enzymatic activity. Higher pectin concentrations increase the viscosity of the medium, agitation being indispensable to promote contact of the enzyme with the substrate. But at low concentrations such as that used in this work, high agitation speeds may hinder the formation of the enzyme-substrate complex. SongpimSONGPIM M, VAITHANOMSAT P and CHUNTRANULUCK S. 2010. Optimization of pectate lyase production from Paenibacillus polymyxa N10 using response surface methodology. Open Biol J 3: 1-7. et al. (2010), working with pectate lyase enzyme, studied the effect of agitation speeds from 150 at 250 rpm and maximum response was obtained with 200 rpm, demonstrating the effect of this variable on the enzymatic activity.

In general, the enzyme activity increases as enzyme concentration increases. However, this increase is limited until the level of substrate saturation, as was demonstrated by MichizoeMICHIZOE J, GOTO M and FURUSAKI S. 2001. Catalytic activity of laccase hosted in reversed micelles. J Biosci Bioeng 92: 67-71. et al. (2001) when they studied the laccase concentration effect from 0 at 15 μM over degradation of o-chlorophenol. The authors observed that after 11 μM, the degradation rate was practically stable.

COMPARISON BETWEEN KINETICS OF ACID AND ENZYMATIC HYDROLYSIS

The kinetics during at acid hydrolysis (Figure 3a) revealed a higher rate of hydrolysis during the first 15 minutes, with a gradual increase up to 4 hours, when a maximum concentration of 6.0 g/L was achieved (which was higher than the 5.3 g/L predicted by the model). Then there was a slight decline until the end of the 5-hour hydrolysis period. The release profiles of the carbohydrate groups were in agreement with the results obtained for RCs, with the concentration of GalA (maximum 3.2 g/L reached in a 4-hour hydrolysis period) exceeding that of NeutralS (maximum concentration 1.8 g/L reached in a 1-hour hydrolysis period).

Figure 3
Release kinetics of reducing compounds during acid hydrolysis of pectin (a) – with 1.0% v/v H2SO4; 100 oC; 1.0% (w/v) pectin; and during enzymatic hydrolysis of pectin (b) - with 10.01 UI/g of enzyme; 230.3 rpm; 1.0% (w/v) of pectin.

Pectins with a high degree of esterification are commonly extracted using hot water (60 at 100 oC) at pH ranges from 1.5 to 3.0 for several hours (KoubalaKOUBALA BB, MBOME LI, KANSCI G, MBIAPO FT, CREPEAU MJ, THIBAULT JF and RALETD MC. 2008. Physicochemical properties of pectins from ambarella peels (Spondias cytherea) obtained using different extraction conditions. Food Chem 106: 1202-1207. et al. 2008). These extraction conditions of pectin could explain the elevated initial rate of free GalA. The resistance to acid hydrolysis of the glycosidic linkages is variable and runs as follows: GalA---GalA > GalA---Rha > Rha---GalA > sugar neutral–sugar neutral, GalA monomers being the last to be released (NovoselskayaNOVOSELSKAYA IL, VOROPAEVA NL, SEMENOVA LN and RASHIDOVA SS. 2000. Trends in the science and applications of pectins. Chem Nat Compd 36: 1-10. et al. 2000). This could explain the gradual release of RCs up to the end of the hydrolysis process. The lower resistance of the glycosidic linkages between neutral sugars are perceived in the behavior curve of the NeutralS that have a higher rate of release in the first 30 min of hydrolysis.

These results are similar to those obtained in earlier studies by Garna et al. (2006), who hydrolysed a highly esterified pectin using different concentrations of sulfuric acid at 100 oC. For all treatments, a gradual release of GalA was observed during the first hours of the process, a decline in the concentration of free GalA being observed after different times for each treatment. According to the authors, the continuation of hydrolysis conditions for long periods led to rates of destruction of free GalA that exceeded the release rates.

In addition to the GalA monomers, the pentoses present in pectin can be degraded to formation of furfural derivatives, which makes the precise determination of sugars impossible (Wikiera et al. 2015, MedinaMEDINA JC, JENSEN GO and NERI JP. 1942. Composição química da casca de ramí em diversas fases do seu desenvolvimento. Bol Tec Div Exp Pesqui Inst Agron 2: 433-447. et al. 1942). The results indicated that formation of these degradation products was low for up to 4 hours of hydrolysis under the conditions employed, so that the hydrolysate could therefore be used in the culture medium.

In order to confirm the results obtained in enzymatic hydrolysis optimization, it was performed an experiment in triplicate using the optimized conditions achieved previously. The RC release kinetics was monitored during 24 hours of enzymatic hydrolysis (Figure 3b). The rate of hydrolysis was faster during the first 30 minutes and remained high for 8 hours. This was succeeded by a gradual hydrolysis for the remaining period of 24 hours, and at the end the RC concentration was 9.34 ± 0.16 g/L (which was greater than the value of 8.85 g/L predicted by the model).

The release profiles of the GalA and NeutralS groups were broadly similar to that of the RCs. There was a greater initial release of the GalA during the first hour of hydrolysis, followed by gradual release for the remaining period of 24 hours and reaching a maximum concentration of 8.4 g/L. The NeutralS group showed a gradual release up to 19 hours, after which the concentration remained constant up to 24 hours of hydrolysis (maximum concentration 1.4 g/L).

Bélafi-BakóBÉLAFI-BAKÓ K, ESZTERLE M, KISS K, NEMESTÓTHY N and GUBICZA L. 2007. HYDROLYSIS of pectin by Aspergillus niger polygalacturonase in a membrane bioreactor. J Food Eng 78: 438-442. et al. (2007) also observed a decrease in the RC and GalA release rate during the course of the hydrolysis process. Those authors utilized different initial concentrations of free GalA and demonstrated that the product of hydrolysis inhibited the enzymatic activity. The neutral sugars, found at lower concentrations in pectin, are probably depolymerized due to the lower resistance of their glycosidic bonds (Novoselskaya et al. 2000), with an acidity of pH 4.0 and a temperature of 50 oC being sufficient to cause their release.

Given an initial pectin concentration of 1.0% (w/v), the average yields of the enzymatic and acid hydrolyses were 93.0% and 60.0%, respectively. The enzymatic method was therefore more efficient for the production of RCs from the hydrolysis of pectin.

CELL GROWTH

The hydrolysates obtained by both methods were diluted for use in the culture medium formulations. Initial concentrations of RCs were different because it was considered the initial substrate concentration in the hydrolysis process (pectin 1.0% w/v). Therefore, it achieved 3.82 and 5.30 g/L RCs in the initial media formulated with the hydrolyzed acid and enzymatic, respectively. The growth of C. necator and parameters kinetics of process were followed as a function of time. Cell concentration, pH, and substrate consumption are shown in Figures 4 and 5.

When the chemical hydrolysate was used (Figure 4a), growth of the microorganism began immediately after cell inoculation and continued for around 10 hours, with μMax of 0.26 h-1. For the culture using the enzymatic hydrolysate (Figure 5a), there was an adaptive phase lasting for around 2 hours, followed by exponential growth up to 12 hours, with μMax of 0.29 h-1. The results can be compared with those obtained by other authors (Table V).

Figure 4
Cell growth and pH (a); and consumption of substrates (b), during culture of C. necator in mineral media containing chemical pectin hydrolysate.
Figure 5
Cell growth and pH (a); and consumption of substrates (b), during culture of C. necator in mineral media containing enzymatic pectin hydrolysate.
TABLE V
Cell growth rates of Cupriavidus necator using Chemical and Enzymatic Hydrolysate, compared with other carbon sources.

The pH remained at around 7.0 throughout the culture period, using both formulations. The consumption of RCs was related to the growth phase, with residual values of 2.8 and 3.7 g/L for media formulated with the hydrolyzed, acid and enzymatic, respectively (Figure 4b and 5b). We see a residual amount of RCs that can’t be metabolized by C. necator, which was also observed by other authors using other carbon sources (BaeiBAEI MS, NAJAFPOUR GD, YOUNESI H, TABANDEH F and EISAZADEH H. 2009. Poly(3-hydroxybutyrate) synthesis by Cupriavidus necator DSMZ 545 utilizing various carbon sources. World Appl Sci J 7: 157-161. et al. 2009, 2011, Locatelli et al. 2011, LagunesLAGUNES FG and WINTERBURN JB. 2016. Effect of limonene on the heterotrophic growth and polyhydroxybutyrate production by Cupriavidus necator H16. Bioresour Technol 221: 336-343. and Winterburn 2016).

This RC residual in both hydrolysates could be related to the presence of oligomers with reducing terminal residues that the microorganism is unable to metabolize. The furanic aldehydes formed by degradation of GalA and sugar affect the microorganism’s metabolism, being toxic for fungus (SzengyelSZENGYEL Z and ZACCHI G. 2000. Effect of acetic acid and furfural on cellulose production of Trichoderma reesei RUT C30. Appl Biochem Biotechnol 89: 31-42. and Zacchi 2000), yeasts (TaherzadehTAHERZADEH MJ, GUSTAFSSON L, NIKLASSON C and LIDÉN G. 1999. Conversion of furfural in aerobic and anaerobic batch fermentation of glucose by Saccharomyces cerevisiae. J Biosci Bioeng 87: 169-174. et al. 1999) and bacteria (ZaldivarZALDIVAR J, MARTINEZ A and INGMAR L. 1999. Effect of selected aldehydes on the growth and fermentation of ethanologenic Escherichia coli. Biotechnol Bioeng 65: 24-33. et al. 1999) and undesirables in culture medium formulations, which could explain the low final cellular concentration using the medium formulated with the hydrolyzed acid.

CONCLUSIONS

The production of RCs from the hydrolysis of pectin was more efficient using an enzymatic method, with a yield 33.0% higher than that achieved using acid hydrolysis. The release profiles of the GalA and NeutralS groups were broadly similar to that of the RCs, to both hydrolysis processes. Moreover, it proves that drastic conditions as the high acid’ concentration can be negative over RC’ production.

The mineral medium formulation containing an enzymatic hydrolysate provided a higher final cell concentration during growth of C. necator, with a specific growth rate that was superior to that obtained using a chemical hydrolysate. In this way, enzymatic hydrolysis can be used in the saccharification of agroindustrial waste pectin, and the hydrolysis product can be applied in the process of C. necator growth, with potential for SCP production or for PHA production.

ACKNOWLEGMENTS

The authors are very grateful for financial support from the Fundação de Amparo à Ciência e Tecnologia do Estado de Pernambuco (FACEPE), for the master’s degree scholarship, and to Professor Nelson Medeiros de Lima Filho, for assistance with the chromatographic analysis.

REFERENCES

  • ADETUNJI LR, ADEKUNLE A, ORSAT V and RAGHAVAN V. 2017. Advances in the pectin production process using novel extraction techniques: A review. Food Hydrocoll 62: 239-250.
  • AKARAONYE E, KESHAVARZ T and ROY I. 2010. Production of polyhydroxyalkanoates: the future green materials of choice. J Chem Technol Biotechnol 85: 732-743.
  • ALBUQUERQUE PBS and MALAFAIA CB. 2018. Perspectives on the production, structural characteristics and potential applications of bioplastics derived from polyhydroxyalkanoates. Int J Biol Macromol 107: 615-625.
  • ARAGÃO GMF, LINDLEY ND, URIBELARREA JL and PAREILLEUX A. 1996. Maintaining a controlled residual growth capacity increases the production of polyhydroxyalkanoate copolymers by A. eutrophus. Biotechnol Lett 18: 937-942.
  • ARAMVASH A, SHAHABI ZA, AGHJEH DS and GHAFARI MD. 2015. Statistical physical and nutrient optimization of bioplastic polyhydroxybutyrate production by Cupriavidus necator. Int J Environ Sci Technol 12: 2307-2316.
  • BAEI MS, NAJAFPOUR GD, YOUNESI H, TABANDEH F and EISAZADEH H. 2009. Poly(3-hydroxybutyrate) synthesis by Cupriavidus necator DSMZ 545 utilizing various carbon sources. World Appl Sci J 7: 157-161.
  • BAEI MS, NAJAFPOUR GD, YOUNESI H, TABANDEH F, ISSAZADEH H and KHODABANDEH M. 2011. Growth kinetic parameters and biosynthesis of polyhydroxybutyrate in Cupriavidus necator DSMZ 545 on selected substrates. Chem Ind Chem Eng Q 17: 1-8.
  • BÉLAFI-BAKÓ K, ESZTERLE M, KISS K, NEMESTÓTHY N and GUBICZA L. 2007. HYDROLYSIS of pectin by Aspergillus niger polygalacturonase in a membrane bioreactor. J Food Eng 78: 438-442.
  • BIERMANN CJ. 1988. Hydrolysis and other cleavages of glycosidic linkages in polysaccharides. Adv Carbohydr Chem Biochem 46: 251-272.
  • CAFFALL KH and MOHNEN D. 2009. The structure, function, and biosynthesis of plant cell wall pectic polysaccharides. Carbohydr Res 344: 1879-1900.
  • CAVALHEIRO JMBT, DE ALMEIDA MCMD, GRANDFILS C and DAFONSECA MMR. 2009. Poly(3-hydroxybutyrate) production by Cupriavidus necator using waste glycerol. Process Biochem 44: 509-515.
  • DALCANTON F, IENCZAK JL, FIORESE ML and ARAGÃO GMF. 2010. Produção de poli(3-hidroxibutirato) por Cupriavidus necator em meio hidrolisado de amido de arroz com suplementação de óleo de soja em diferentes temperaturas. Quím Nov 33: 552-556.
  • De RUITER GA, SCHOLS HA, VORAGEN AGJ and ROMBOUTS FM. 1992. Carbohydrate analysis of water-soluble uronic acid containing polysaccharides with high-performance anion-exchange chromatography using methanolysis combined with TFA hydrolysis is superior to four other methods. Anal Biochem 207: 176-185.
  • FIGUEIREDO TVB, CAMPOS MI, SOUZA LS, SILVA JR and DRUZIANA JI. 2014. Production and characterization of polyhydroxyalkanoates obtained by fermentation of crude glycerin from biodiesel. Quim Nov 37: 1111-1117.
  • GARNA H, MABON N, NOTT K, WATHELET B and PAQUOT M. 2006. Kinetic of the hydrolysis of pectin galacturonic acid chains and quantification by ionic chromatography. Food Chem 96: 477-484.
  • JÖRNEDING HJ, BACIO IE, BERENSMEYER S and BUCHHOLZ K. 2002. Gewinnung von galacturonsäure aus zellwandbestandteilen der rübenschnitzel. Zuckerind 127: 845-853.
  • KLEIN H and LEUBOLT R. 1993. Ion-exchange high-performance liquid chromatography in the brewing industry. J Chromatogr 640: 259-270.
  • KISS K, CSERJÉSI P, NEMESTÓTHY N, GUBICZA L and BÉLAFI-BAKÓ K. 2008. Kinetic study on hydrolysis of various pectins by Aspergillus niger polygalacturonase. Hung J Ind Chem 36: 55-58.
  • KOHLI P and GUPTA R. 2015. Alkaline pectinases: A review. Biocatal Agric Biotechnol 4: 279-285.
  • KOUBALA BB, MBOME LI, KANSCI G, MBIAPO FT, CREPEAU MJ, THIBAULT JF and RALETD MC. 2008. Physicochemical properties of pectins from ambarella peels (Spondias cytherea) obtained using different extraction conditions. Food Chem 106: 1202-1207.
  • KUNASUNDARI B, MURUGAIYAH V, KAUR G, MAUER FHJ and SUDESH K. 2013. Revisiting the Single Cell Protein Application of Cupriavidus necator H16 and Recovering Bioplastic Granules Simultaneously. Plos One 8: 1-15.
  • LAGUNES FG and WINTERBURN JB. 2016. Effect of limonene on the heterotrophic growth and polyhydroxybutyrate production by Cupriavidus necator H16. Bioresour Technol 221: 336-343.
  • LEITÃO MCA, ALARCÃO-SILVA ML, JANUÁRIO MIN and AZINHEIRA HG. 1995. Galacturonic acid in pectic substances of sunflower head residues: quantitative determination by HPLC. Carbohydr Polym 26: 165-169.
  • LIM J, YOO J, KO S and LEE S. 2012. Extraction and characterization of pectin from Yuza (Citrus junos) pomace: A comparison of conventional-chemical and combined physical-enzymatic extractions. Food Hydrocoll 29: 160-165.
  • LOCATELLI GO, SILVA GD, FINKLER L and FINKLER CLL. 2011. Acid hydrolysis of pectin for cell growth of Cupriavidus necator. Biotechnol 1: 1-8.
  • MARANGONI C, FURIGO JRA and ARAGÃO GMF. 2001. The influence of substrate source on the growth of Ralstonia eutropha, aiming at the production of polyhydroxyalkanoates. Brazilian J Chem Eng 18: 175-180.
  • MAY CD. 2000. Pectins. In: Phillips GO and Williams PA (Eds), Handbook of Hydrocolloids, New York: CRC Press, p. 165-185.
  • MEDINA JC, JENSEN GO and NERI JP. 1942. Composição química da casca de ramí em diversas fases do seu desenvolvimento. Bol Tec Div Exp Pesqui Inst Agron 2: 433-447.
  • MICHIZOE J, GOTO M and FURUSAKI S. 2001. Catalytic activity of laccase hosted in reversed micelles. J Biosci Bioeng 92: 67-71.
  • MILLER GL. 1959. Use of dinitrosalicylic acid reagent for determination of reducing sugar. Anal Chem 31: 426-428.
  • MIN B, LIM J, KO S, LEE KG, LEE SH and LEE S. 2011. Environmentally friendly preparation of pectins from agricultural byproducts and their structural / rheological characterization. Bioresour Technol 102: 3855-3860.
  • MUZZARELLI RAA, BOUDRANT J, MEYER D, MANNO N, DEMARCHIS M and PAOLETTI MG. 2012. Current views on fungal chitin/chitosan, human chitinases, food preservation, glucans, pectins and inulin: A tribute to Henri Braconnot, precursor of the carbohydrate polymers science, on the chitin bicentennial. Carbohydr Polym 87: 995-1012.
  • NOVOSELSKAYA IL, VOROPAEVA NL, SEMENOVA LN and RASHIDOVA SS. 2000. Trends in the science and applications of pectins. Chem Nat Compd 36: 1-10.
  • PINTO GAS, BRITO ES, SILVA FLH, SANTOS SFM and MACEDO GR. 2006. Fermentação em estado sólido: Uma alternativa para o aproveitamento e valorização de resíduos agroindustriais. Rev Quim Ind 74: 17-20.
  • RAMSAY BA, LOMALIZA K, CHAVARIE C, DUBE B, BATAILLE P and RAMSAY JA. 1990. Production of poly-(β-hydroxybutyric-co-β-hydroxyvaleric) acids. Appl Environ Microbiol 56: 2093-2098.
  • SONGPIM M, VAITHANOMSAT P and CHUNTRANULUCK S. 2010. Optimization of pectate lyase production from Paenibacillus polymyxa N10 using response surface methodology. Open Biol J 3: 1-7.
  • SZENGYEL Z and ZACCHI G. 2000. Effect of acetic acid and furfural on cellulose production of Trichoderma reesei RUT C30. Appl Biochem Biotechnol 89: 31-42.
  • TAHERZADEH MJ, GUSTAFSSON L, NIKLASSON C and LIDÉN G. 1999. Conversion of furfural in aerobic and anaerobic batch fermentation of glucose by Saccharomyces cerevisiae. J Biosci Bioeng 87: 169-174.
  • WANG Y, YIN J and CHEN GQ. 2014. Polyhydroxyalkanoates, challenges and opportunities. Curr Opin Biotechnol 30: 59-65.
  • WENZEL GE. 2001. Bioquímica Experimental dos Alimentos, 1a ed., Unisinos, Rio Grande do Sul, 213 p.
  • WIKIERA A, MIKA M, STARZYNSKA-JANISZEWSKA A and STODOLAK B. 2015. Development of complete hydrolysis of pectins from apple pomace. Food Chem 172: 675-680.
  • YOUSUF RG and WINTERBURN JB. 2016. Date seed characterisation, substrate extraction and process modelling for the production of polyhydroxybutyrate by Cupriavidus necator. Bioresour Technol 222: 242-251.
  • ZALDIVAR J, MARTINEZ A and INGMAR L. 1999. Effect of selected aldehydes on the growth and fermentation of ethanologenic Escherichia coli. Biotechnol Bioeng 65: 24-33.

Publication Dates

  • Publication in this collection
    25 Apr 2019
  • Date of issue
    2019

History

  • Received
    12 Apr 2018
  • Accepted
    24 July 2018
Academia Brasileira de Ciências Rua Anfilófio de Carvalho, 29, 3º andar, 20030-060 Rio de Janeiro RJ Brasil, Tel: +55 21 3907-8100 - Rio de Janeiro - RJ - Brazil
E-mail: aabc@abc.org.br