Next Article in Journal
Physicochemical and Microstructural Characterization of Whey Protein Films Formed with Oxidized Ferulic/Tannic Acids
Next Article in Special Issue
Impact of Oral Microbiota on Flavor Perception: From Food Processing to In-Mouth Metabolization
Previous Article in Journal
Backhousia citriodora F. Muell. (Lemon Myrtle), an Unrivalled Source of Citral
Previous Article in Special Issue
Contribution of S. xylosus and L. sakei ssp. carnosus Fermentation to the Aroma of Lupin Protein Isolates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biochemical Properties Affecting the Nutritional Quality, Safety, and Aroma of Dry-Cured Products Manufactured from Meat of Rare Native Pig Breeds

by
Ewelina Węsierska
1,*,
Joanna Sobolewska-Zielińska
2,
Małgorzata Pasternak
1,
Katarzyna Niemczyńska-Wróbel
1,
Robert Gąsior
3,
Krzysztof Wojtycza
3,
Henryk Pustkowiak
4,
Iwona Duda
1 and
Władysław Migdał
1
1
Department of Animal Product Technology, University of Agriculture in Krakow, Balicka 122, 30-149 Krakow, Poland
2
Department of Food Analysis and Evaluation of Food Quality, University of Agriculture in Krakow, Balicka 122, 30-149 Krakow, Poland
3
Central Laboratory, National Research Institute of Animal Production, Balice, 32-083 Krakow, Poland
4
Department of Cattle Breeding, University of Agriculture in Krakow, Al. Mickiewicza 24/28, 30-059 Krakow, Poland
*
Author to whom correspondence should be addressed.
Foods 2021, 10(7), 1597; https://doi.org/10.3390/foods10071597
Submission received: 18 May 2021 / Revised: 29 June 2021 / Accepted: 7 July 2021 / Published: 9 July 2021
(This article belongs to the Special Issue Flavour Research in Food Fermentations)

Abstract

:
The aim of study was to compare the biochemical properties affecting the nutritional quality, safety, and aroma of dry-cured products manufactured from valuable meat of rare native pig breeds: Pulawska (Pul) and Zlotnicka Spotted (ZS). The count of lactic acid bacteria (4.4 log cfu/g) and the release of palmitic (23.1% and 25.9%), oleic (44.1% and 42.2%), and linoleic acids (8.3% and 7.8%), as well as arginine (30.0 and 44.3 mg/kg), histidine (25.8 and 20.6 mg/kg), and lysine (26.8–22.9 mg/kg), shaped the final pH (5.3 and 5.4) in Pul and ZS products during the 4 week maturing, respectively. Lastly, Pul and ZS meat differed in the proportion of decanoic, lauric, stearic, arachidic, and conjugated linoleic acids. The high content of putrescine (23.7 mg/kg), cadaverine (54.3 mg/kg), and tyramine (57.2 mg/kg), as well as a twofold greater share of histamine (163.2 mg/kg) and tryptamine (9.1 mg/kg), indicated a more advanced decarboxylation of ZS meat. Volatile compounds differentiating Pul and ZS meat were primarily hexanal, 3-hydroxybutan-2-one, phenylacetalaldehyde, 2,3-dimethyl-2-cyclopenten-1-one, 2-cyclopenten-1-one, and 3-methyl- and 2-cyclopenten-1-one. Most marked volatile compounds were obtained as a result of microbial activity (acetic acid, 3-methylbutan-1-ol, ethanol, acetone, and 3-hydroxybutan-2-one), advanced lipid oxidation, and decomposition of secondary oxidation products (hexanal, phenylacetaldehyde, and 2-cyclopenten-1-one).

Graphical Abstract

1. Introduction

The pork of native breeds Pulawska (Pul) and Zlotnicka Spotted (ZS), due to the appropriate proportion of water (69.8–74.0% and 69.2–73.6%, respectively) to other basic chemical components and sufficient drip loss (1.7–4.7% and 2.4–3.4%, respectively), is beneficial for manufacturing of regional products of a unique culinary value. For comparison, the commercially used pigs give incomparable drip loss values of 3.0–4.3% [1,2,3,4,5,6]. The studies of Kapelański et al. [7] and Grześkowiak et al. [8] revealed PSE (pale, soft, exudative) and DFD (dark, firm, dry) defects only for single Pul and ZS carcasses. These observations confirmed the correct course of post-slaughter maturation, suitably low pH values, and technological usefulness of the meat. The quality of Pul and ZS meat does not differ significantly in terms of pH, brightness, and cutting force, but there are some differences in the chemical composition [9]. Slightly different marbling creates not only an individual sensory appearance and attractiveness of slices on the cross-section, but also a juiciness and a distinct flavor of Pul and ZS meat [7,8]. The high content of protein shapes the nutritional value and texture different to industrial breed meat [10]. Meat of the Pul breed is characterized mainly by special organoleptic value and is suitable for the production of regional slow food. The high quality of ZS pork allows it to be used for the production of cooked and baked hams and sirloins [1,7,9,11,12,13,14]. The meat of Pulawska and Zlotnicka White and Spotted breeds of pigs is a good-quality raw material used for production of traditional and regional smoked meat products gaining high sensory scores and good recognition among consumers [15,16]. For this reason, the production of dry-cured, long-maturing meat products is technologically attractive to better understand the raw material and the transformations that it undergoes. The European Food Information Resource Network (EuroFIR) indicates that traditional products have specific features that make it possible to distinguish them from other similar products. Traditional products of animal origin should have a traditional recipe and a traditional method of production, while they should be made from traditional raw materials, preferably using native animal breeds. A similar direction of meat production from native breeds can be found in southern European Union countries. The free-range system increases the value of animal products due to the influence of outdoor rearing on the chemical, physical, and organoleptic characteristics of the product. Traditional products made from pork of native breeds are highly appreciated: Spanish “Ibérico” ham, Tuscan “Cinta senese”, Basque ham, and the Corsican prisuttu are produced from Iberian, Cinta Senese, Pie Noir du pays Basque, and Corsican Black or Spotted native pig meat, respectively. Long-term maturing is done in appropriate climatic conditions, at a temperature of 15–18 °C and a relative humidity of 85–90%, lowered to 70–80% at the further stage of maturing. The temperature of the curing smoke at the time of cold smoking which lasts for several days should not be higher than 18–25 °C [17]. Long-term maturing smoked products constitute an important segment of regional products in Italy, Spain, Portugal, France, and recently even the Czech Republic.
The aim of the study was to compare the progress of 4 week maturation of dry-cured ham made from the meat of two native breeds of pigs (Pulawska and Zlotnicka Spotted), spontaneously fermented using the same method. The progress of drying was assessed on the basis of changes in the chemical composition (water, protein, fat, ash, salt content), pH, and water activity, as well as the succession of acidifying and denitrifying microflora. The advancement of the proteolysis, amino-acid decarboxylation, lipolysis, and fat oxidation was studied by measuring the share of amino acids, amine groups, biogenic amines, fatty-acid content, and TBA index, respectively. Volatile compounds were analyzed to illustrate the advancement of microbial fermentation, amino-acid and fatty-acid catabolism, and lipid oxidation. Since meat from the native breeds is an excellent raw material for the production of slow food, the analyses planned in this study were carried out during the 4 week ripening of dry-cured hams, low-processed and produced using the same traditional method with the complete absence of advanced technologies, including heat treatment.

2. Materials and Methods

2.1. Dry-Cured Product Manufacturing

The raw material was the topside part of ham muscles after the group of pelvic bones group was removed, taken from pigs of native breeds Pulawska (Pul) and Zlotnicka Spotted (ZS). Thirty fatteners of each of the discussed breeds were chosen. The pigs were fed ad libitum with a complete feeding mixture to obtain a body weight of 125–130 kg. Three times, after reaching the slaughter mass, pigs were slaughtered (the cuts for production were randomly selected from a pool of 30 carcasses). The meat was obtained from the same litter of pigs (half-siblings). All activities related to slaughter and post-slaughter processing were carried out in an industrial slaughterhouse in accordance with applicable regulations. The weight of chilled ham after boning varied between 1.2 and 1.3 kg. To obtain a salt content of no more than 4.5% in the ready-to-eat products, salt was added during dry curing in the amount of 2% of non-iodized salt and 1% of curing salt (min. 98.4% NaCl, 0.6% NaNO2, Anna, Krakow, Poland) related to the weight of the raw meat. Additionally, 0.3% sugar and spices (0.9% juniper berries, 0.4% allspice, 0.4% pepper, 0.4% garlic) were added in relation to the meat weight. The ham was placed in containers for curing in dry (2 weeks) and then wet conditions (1 week) at 4–7 °C. The ratio of the amount of the brine to the meat was 2:3. The 8° brine contained 87 g of NaCl per 1000 mL of water. The cooled cure was slightly boiled beforehand to eliminate any accidental microflora. Ham was smoked cold three times (15–25 °C) with alder-beech chips in daily intervals (8 h/day). Ripening at a temperature of 10–12 °C and 85–90% relative humidity (first week) and 85% (subsequent time of maturing) was carried out in a climate chamber KK350 TOP+ (POL-EKO) for 4 weeks. The slaughter was carried out three times. The obtained carcasses were used each time for three batches of ham production. Three whole dry-cured hams (three Pul and three ZS products) as replications in each batch and three samples from each ham were analyzed.

2.2. Sampling

Samples of the raw material were collected just after dry salting and the smoking procedures (ripening period: 0). The cuts were also sampled at different times throughout the maturing (after second and four weeks). The material for analysis was cut into 1.5 cm thick slices. The first one (external) was rejected because of its dried surface and influential water loss. After separating, the samples for the basic chemical composition (water, protein, fat, ash, salt), pH, water activity, and the total count of microorganisms, as well as the counts of acidifying and denitrifying microflora, were estimated as the process of maturing progressed. The contents of free amino acids, free amine groups, and biogenic amines, as well as the profiles of fatty acids and volatile compounds, were determined only in the dry-cured ham ready for consumption, after 4 weeks of maturing. All meat samples for analysis were transported to the appropriate laboratories under refrigerated conditions. Analyses were done immediately after delivery. The test samples were prepared in accordance with the appropriate standards.

2.3. Analysis

2.3.1. Basic Chemical Composition and Physicochemical and Microbiological Characteristics

The basic chemical composition was evaluated according to the Polish and European recognized standards. The moisture and ash contents were determined by drying samples to their stable weight [18]. The protein content was determined with the use of the Kjeldahl method with the set-type 322 (Büchi, Flawil, Switzerland) using a 6.25 factor [19]. Free fat was evaluated with the use of the Soxhlet method with ethyl ether extraction [20] and the chloride content with the use of Volhard’s method [21]. The energy value of meats was determined by the Atwater specific factors of Food and Agriculture Organization of the United Nations (2003). The pH was measured with pH-meter type CP-411 and electrode type PP-3 (Elmetron, Zabrze, Poland) in water homogenate (meat–water ratio, 1:3) and water activity was determined with the Lab Master (Novasina, Lachen, Switzerland), following the producer manual. Analyses for bacteriological examination were done for the following: the total count of microorganisms in mesophilic conditions (Standard Methods Agar, Biomérieux, 30 °C/72 h) [22], acidifying microflora (MRS Agar, Biomérieux, acetic acid used for pH reducing to 5.4, 30 °C/24–48 h, anaerobic chamber with a 20% CO2 enriched atmosphere (Sheldon Manufacturing Inc., Cornelius, OR, USA) [23], and denitrifying microflora (Baird Parker Agar Base, yolk emulsion and sodium tellurite, Biomérieux; coagulase-negative cocci classified on the basis of bound coagulase and free coagulase activity, 37 °C/24 h) [24].

2.3.2. Proteolytic Changes and Biogenic Amine Content

Free amino groups were evaluated in water and in phosphotungstic acid [25]. Water-soluble nitrogen (WSN) was determined in the water-soluble fraction (WSF) of the sample. The contents of free amino-acid groups in water (WSF) and the phosphotungstic acid (PTA)-soluble fractions were measured using a 2,4,6-trinitrobenzenesulfonic acid (TNBS) (Sigma-Aldrich, Saint Louis, MI, USA). A phosphotungstic acid soluble fraction of meat sample was prepared from WSF. The WNS fraction was used to determine the free amino acids as follows: 5 mL of the water-soluble protein fraction was added to 5 mL of 40% trichloroacetic acid (TCA) and stirred on a Vortex RS-va10 (Phoenix Instrument, Garbsen, Germany) for 10 min, before centrifuging at 12,000× g for 10 min in a precooled centrifuge MPW-352R, REF (MPW Med. Instruments, Warszawa, Poland) at 4 °C. Then, 2 mL of solution was removed and dried under nitrogen. The obtained residue was dissolved in 1 mL of 20 mM HCl. The chromatographic separation (HPLC) of free amino acids was performed with Dionex UltiMate 3000 (Thermo Scientific, Waltham, MA, USA). Separation was carried in a Nova-Pak reverse-phase C18 column, 4 µm particle size, 150 × 3.9 mm (Waters, Milford, MA, USA), thermostated at 37 °C. The two solvent reservoirs contained the following eluents: (A) acetate–phosphate buffer pH = 5.2 and (B) a mixture of acetonitrile and water (60:40, v/v) according to the Waters procedure. The elution program consisted of a gradient system with a flow rate of 1 mL·min−1 (100% A for 0.5 min, 98% A and 2% B for 14.5 min, 93% A and 7% B to 19 min, 90% A and 10% B to 32 min, 67% A and 33% B to 33 min, 100% B to 37 min, and 100% A to 64 min). Fluorometric detection was carried out using excitation and emission wavelengths of 250 nm and 395 nm, respectively. Analytical kit ACCQ Tag (Waters, USA) was used as a reference.
The biogenic amine analysis was performed using HPLC according to the following method: a 10 g sample of meat was mixed with 15 mL of 6% trichloroacetic acid (TCA), homogenized for 2 min (1000 rpm, Vortex), and centrifuged at 14,000× g for 20 min at 4 °C (MPW Med. Instruments, Warszawa, Poland). The obtained supernatant was collected and filtered through Whatman No. 1 filter paper, and the residue was reextracted with 15 mL of fresh of 6% TCA. The volume of the combined supernatants was completed to 50 mL by 6% TCA. The derivatization process was performed by mixing 1 mL of extract with 1 mL of dansyl chloride acetone solution (10 mg·mL−1) and 0.5 mL of saturated NaHCO3 solution. Incubation was carried out at 40 °C for 60 min with occasional shaking in a thermoblock (Dry Block Heater, IKA, Staufen, Germany). The biogenic amines were extracted twice using 1 mL of diethyl ether for 10 min (22 °C). The extracts were dried in a stream of nitrogen, and the residue was dissolved in 1 mL of acetonitrile. The solution standards of biogenic amines were prepared by mixing 1 mL of each free base standard solution, containing 0.1 mg·mL−1 of each biogenic amine with 5 mL of dansyl chloride acetone solution. The derivatization and extraction procedure was performed in the same manner as the samples. The chromatographic separation of free amino acids was performed using Dionex UltiMate 3000 HPLC apparatus (Thermo Scientific, Waltham, MA, USA). The separation was carried in a Nova-Pak reverse-phase C18 column, 4 µm particle size, 150 × 3.9 mm (Waters, Milford, MA, USA), thermostated at 30 °C. The two solvent reservoirs contained the following eluents: (A) acetonitrile and (B) HPLC-grade water at a flow rate of 0.8 mL·min−1. The elution program was: 65% A and 35% B for 1 min, increasing to 80% A and 20% B for 9 min, increasing to 90% A and 10% for 2 min, increasing to 95% A and 5% B for 4 min, holding for 7 min, and returning to 65% A and 35% B, before holding for 5 min. Fluorometric detection was carried out using excitation and emission wavelengths of 340 nm and 530 nm, respectively. The solution of standards of eight biogenic amines was prepared in 0.1 M HCl in a concentration of 1 mg·mL−1 of every amine; each of the following reference substances was dissolved in 10 mL of 0.1 M HCl: tyramine hydrochloride (Sigma-Aldrich, Buchs, Switzerland)—12.7 mg, tryptamine hydrochloride (Sigma-Aldrich, Saint Louis, MI, USA)—12.3 mg, histamine dihydrochloride (Sigma-Aldrich, Beijing, China)—16.6 mg, putrescine dihydrochloride (Sigma-Aldrich, Buchs, Switzerland)—18.3 mg, cadaverine dihydrochloride (Sigma-Aldrich, Buchs, Switzerland)—17.1 mg, spermine tetrahydrochloride (Sigma-Aldrich, Buchs, Switzerland)—17.2 mg, spermidine trihydrochloride (Sigma-Aldrich, Buchs, Switzerland)—17.5 mg, and 2-phenylethylamine hydrochloride (Sigma-Aldrich, Tokyo, Japan)—13.0 mg. Prepared solution standards were then 100-fold diluted. All reagents used were HPLC-grade (Sigma-Aldrich, Saint Louis, MI, USA).

2.3.3. Fatty-Acid Profile and Oxidative Changes

The fatty-acid profile was determined by gas chromatography from the fat extracted from the samples of ready-to-eat products. Lipids were extracted from samples with a chloroform–methanol (2:1) mixture, according to the method of Folch, Lees, and Sloane [26]. Trace GC Ultra (Thermo Electron Corporation, Rodano, Italy) with a column Supelcowax 10 (Sigma-Aldrich, Saint Louis, MI, USA) 30 m × 0.25 mm × 0.25 μm was used for the determination. The analysis was carried out under the following conditions: helium carrier gas 1 mL·min−1, split flow 10 mL·min−1, injector temperature 220 °C, detector temperature 250 °C, and starting temperature of the column (first 3 min) 160 °C, increased by 3 °C min−1 to 210 °C and maintained for 25 min. Individual fatty-acid methyl esters were identified by comparison to the standard mixture (Supelco 37 Component FAME Mix, Sigma-Aldrich, USA). The thiobarbituric acid (TBA) index was determined by the spectrophotometric method in a Helios γ spectrophotometer (Thermo Electron Corporation, Beverly, MA, USA) and expressed as mg of malondialdehyde (MDA) in kg of meat. Furthermore, 10 g of sample was homogenized with 34.25 mL of cold 4% perchloric acid and 0.75 mL of alcoholic butylated hydroxytoluene solution (4000 rpm for 2 min in a precooled centrifuge (MPW Med. Instruments, Warszawa, Poland)). Then, 5 mL of the filtrate was combined with 5 mL of 0.02 M TBA and heated for 1 h. Absorbance was read at 530 nm. The blank was 5 mL of 4% perchloric acid solution and 5 mL of TBA [27].

2.3.4. Volatile Compound Content

Volatile compounds were extracted and analyzed using a gas chromatograph–mass spectrometer GCMS-QP 2010 Plus (Shimadzu, Duisburg, Germany), with a 50/30 μm DVB/CAR/PDMS fiber (Supelco, Bellefonte, PA, USA) and Zebron ZB-5MSi and ZB-Wax columns 30 m × 0.25 mm × 0.25 μm (Zebron, Phenomenex®, Torrance, CA, USA) [28]. The identification was done using mass spectral libraries (NIST08, NIST08s, and FF NSC1.3) and retention indices RI, on the basis of analyses of n-paraffins (Supelco, Merck Group, Poznan, Poland), compared with values from National Institute of Standards and Technology. Some of the compounds were verified by standards (S); for quantitation, methyl caproate (internal standard) was used (Sigma-Aldrich, Saint Louis, MI, USA).

2.3.5. Statistical Analysis

The statistical analysis was performed using the Statistica software for Windows, version 13.3 (Statsoft, Krakow, Polska). The effect of the ripening time on the chemical properties and microbiological quality was tested using one-factor (ripening period: 0, 2, 4 weeks) and two-factor (ripening period: 0, 2, 4 weeks; breed: Pul, ZS) analyses of variance (ANOVA with fixed and orthogonal factors). Duncan post hoc tests were used to compare the means at p < 0.05.

3. Results and Discussion

3.1. Basic Chemical Composition, and Physicochemical and Microbiological Characteristics

Despite the removal of the external fat cover of different thickness for both breeds during the boning, Pul and ZS dry-cured products differed in their basic chemical composition. To shorten the presentation, the results are presented in Table 1 as arithmetic means obtained after a two-factor analysis of variance. Pul products had a lower protein (18.8%) and a higher fat content (5.0%) compared to ZS products (22.6% and 4.5%, respectively). According to the differences in the intramuscular fat content of Pul and ZS products, the energy values were calculated as 157.8 and 149.6 kcal, respectively. The basic chemical composition and physicochemical parameters of Pul and ZS products changed significantly over the 4 week maturing. The loss of water from 74.3% to 69.1% (p < 0.05) resulted in an increase in total protein (from 18.2% to 20.8%) (p < 0.05) and fat content (from 3.4% to 6.4%) (p < 0.05). Changes in the basic chemical composition are of great importance for the organoleptic quality and nutritional value of dry-cured products manufactured by both traditional and industrial methods. The share of water, protein, fat, and salt determines the quality properties such as water-holding capacity, juiciness, texture, color, and overall taste during maturing, as confirmed by [15,29,30,31]. The quantitative and qualitative composition of the microflora of raw sausages varies and depends on accidental contamination of the meat. In the initial period of maturation, there is a cenobiotic exchange of microflora from accidental to technologically desirable, i.e., acidifying, denitrifying, and aromatizing. During the initial maturation and cold smoking, the dominant microflora were cocci and rod-shaped bacteria, which lowered the pH and intensified the color of the meat. In the further stage of maturation and during the post-production maturation, staphylococci and micrococci shaped the flavor profile of the meat. Maturation time significantly affected the analyzed total count of microorganisms (p < 0.05), technologically desirable microflora (p < 0.05) (acidifying microflora (lactic acid bacteria) and denitrifying microflora (coagulase-negative cocci)), and the final pH of Pul and ZS products (p < 0.05) (Figure 1). Despite the higher average salt content (1.9%) and the lower pH (5.5), the total count of microorganisms (4.3 log cfu/g) and denitrifying bacteria (2.7 log cfu/g) of Pul was more than that of ZS products. The increased salinity and acidity of dry-cured meat products as a result of maturing and a reduction in the water activity does not inhibit the development of halophilic denitrifying and proteolytic microflora [32,33,34,35]. Staphylococcaceae and Micrococcaceae support the process of color, aroma, and safety formation [29]. An increase in the cocci population observed after the second week of analyzed maturing corresponded to a scenario whereby the fermentation of saccharides (the development of acidifying bacteria) acted most intensively during the first 3–5 days, while the proper development of denitrifying microflora occurred a few weeks later, depending on the type of meat product [36,37].

3.2. Proteolytic Changes and Biogenic Amines Content

The ready-to-eat Pul and ZS dry-cured products differed significantly in the content of amino acids released by proteolysis, especially alanine (47.4 and 30.4 mg/kg), glutamine (44.8 and 23.2 mg/kg), threonine (43.7 and 31.8 mg/kg), and leucine (40.9 and 32.8 mg/kg) (Table 2). The amount of free amino groups dissolved in water determined the majority of typical proteolysis products (0.6 and 0.7 g/100 g (p < 0.05), respectively). Amino groups dissolved in PTA (0.06 g/100 g) in Pul and ZS products proved the presence of the low-molecular-weight fraction of singular amino acids, as well as di- and tripeptides [25,31]. Similar changes in the proteolysis activity and the accumulation of free amino acids during ripening were described in fermented sausages [38], dry-cured ham [39], and Chinese traditional dry-cured loin [40]. Changes in the profiles of metabolites originating from protein degradation affect the sensory quality and safety of dry-cured products during ripening and are still the focus of science [15]. The high proportion of putrescine, cadaverine, and tyramine in ZS (23.7, 54.3, and 57.2 mg/kg) compared to Pul products (4.0, 28.5, and 11.1 mg/kg) confirmed the differences in proteolysis and decarboxylation of ornithine, lysine, and tyrosine, respectively (p < 0.05) (Table 3). A nearly twofold greater share of histamine and tryptamine (163.2 and 9.1 mg/kg, respectively (p < 0.05)) was determined in the ZS products as the result of decarboxylation of histidine and tryptophan. The production of biogenic amines confirmed the metabolic activity of the microflora [41]. The ratio of protein to fat and a slightly different water content could create an environment more or less conducive to the decarboxylation of amino acids by bacteria. The contents of cadaverine, histamine, and tyramine in dry-cured pork shoulder ripened for 3 months remained at a lower level of 7.5, 6.5, and 21.9 mg/kg, respectively [36]. Manufacturing of dry-cured ham is a safe process, although this meat product can occasionally support the accumulation of amino-acid decarboxylation products. In fact, a high proteolytic activity during maturing favors the decarboxylase activity of many microorganisms. So far, no cases of poisoning with biogenic amines after ingestion of dry-cured ham have been confirmed in the literature. The level of amines is not limited by food safety criteria for maturing meat products. The study of Landeta et al. [42] revealed that the production of biogenic amines was not a widely distributed property among staphylococci. Only 3.6% strains of coagulase-negative staphylococci isolated during industrial Spanish dry-cured ham processes could produce some biogenic amines, including histamine, putrescine, cadaverine, and tyramine.
The levels of these amines showed a time-dependent increase during aging (decrease in moisture and water activity, along with a minimal increase in pH and non-protein nitrogen and free amino-acid content), especially in ZS hams, in line with the studies of Lorenzo et al. [43], Martuscelli et al. [44], Stadnik and Dolatowski [45], and Virgili et al. [46].

3.3. Fatty-Acid Profile, Oxidative Changes, and Volatile Compound Content

The dry-cured meat products of Pul and ZS breeds differed in the proportion of selected saturated fatty acids (SFAs): decanoic (C10;0), lauric (C12;0), stearic (C:18;0), and arachidic (C20;0) (p < 0.05) (Table 4). The sum of SFAs in ready-to-eat Pul and ZS products was determined at 34.9% and 38.5%, respectively. Only the content of heptadecenoic acid (C17;1) (p < 0.05) differed in Pul and ZS products, while the sum of MUFAs in both meats amounted to 54.1% and 50.9%, respectively. There were no significant differences in the sum of ω-6 acids in Pul and ZS products. The only acid differentiating Pul and ZS in the ω-3 group was CLA (p < 0.05). In both cases, the sum of ω-3 acids was 0.8%. The quantitative and qualitative composition of fatty acids depends on the composition of the feed mixtures. Pul and ZS pigs were fed the same feed, but their behavior and physiological conditions for fat storage are different. This can change the proportion of fatty acids in meat products. Debrecéni et al. [9], Cebulska [14], and Migdał et al. [15], indicated statistically significant differences in the contents of palmitic acid, oleic acid, LA, ALA, GLA, AA, and DHA in Pul and ZS meat. Probably due to the antioxidative activity of wood pyrolysis, spices, and lactic acid microflora, the TBA index was determined on a low level of 0.3 mg/kg (Pul) and 0.5 mg/kg (ZS) (p < 0.05). Low TBA values may also indicate a very good physicochemical and organoleptic quality of the fat, as well as proper cooling of the carcasses after slaughter [37]. The lipolysis directly resulted in the formation of aroma compounds by liberating fatty acids [33]. MDA enables further transformations such as condensation, reactions with released amino acids, and decomposition via the following catalysts: Pediococcus acidilactici, Lactobacillus plantarum, and Staphylococcus carnosus [29]. Spaziani et al. [32], Ravyts et al. [34] and Berardo et al. [35] indicated aldehydes, ketones, and alcohols as secondary products of hydrolytic rancidity and autooxidation of UFAs. Some of them could be the products of microbial metabolism, including carbohydrate fermentation, as well as FAA catabolism or esterification. Longer chains of fatty acids (C14–C18) could become, along with products of proteolysis, the precursors of selected VOCs. According to Olivares et al. [29] and Węsierska et al. [37], the substances obtained in the enzymatic (20%) and nonenzymatic oxidation of UFAs (80%) become the basis for aroma precursors. Volatile compounds developed from autooxidation of lipids give a rancid profile. However, there was no high content of compounds that would adversely change the sensory profile of the Pul and ZS products. The cause could have been the antioxidative activity of phenols supplied during smoking and absorbed by the surface.
There were 48 volatile compounds (VCs) determined by HS-SPME–GC/MS (Table 5). Substances with the largest amount were obtained through biochemical and microbiological transformations of meat tissues (16 VCs) or came from spices (25 VCs). Some of them were the products of pyrolysis of wood (seven VCs). The compounds that differentiated Pul and ZS products (p < 0.05) were primarily hexanal, 3-hydroxybutan-2-one (acetoin), phenylacetalaldehyde, 2,3-dimethyl-2-cyclopenten-1-one, 2-cyclopenten-1-one, 3-methyl-, 2-cyclopenten-1-one, and 2-methyl- and acetic acid, obtained as a result of microbial activity (fermentation) and lipid oxidation. The compounds that developed the meat aroma in all meats were hexanal, 3-hydroxybutan-2-one, 3-methylbutanal, and 2-methylcyclopentanone. The main sources of sulfur compounds and terpenes were garlic, pepper, and pyrolysis of lignin with their terpy, herbaceous, woody nuances. Some of them occurred in large quantities in Pul products perhaps due to the lower content of protein and higher fat content. The examples of compounds of spices, as well as of pyrolysis origin, that differentiated Pul and ZS products were 2-furanmethanol, 2-furanmethanol, 5-methyl, eugenol, o-xylene, caryophyllene, cresol, diallyl disulfide, camphor, linalool, guaiacol, γ-terpinene, limonene, and 3-carene. The presence of hexanal (11.3 and 42.5 ng/g), phenylacetaldehyde (20.5 and 8.8 ng/g), 2-cyclopenten-1-one (7.2 and 5.9 ng/g), 1-hydroxypropan-2-one (5.9 and 4.9 ng/g), and 2-methylcyclopentanone (11.7 and 11.1 ng/g) could indicate an increased activity of secondary fat oxidation products despite the small values of the TBA index (0.3 and 0.5 mg/kg) in Pul and ZS products, respectively. The amounts of products of saccharide fermentation such as acetic acid, 3-methylbutan-1-ol, ethanol, acetone, and 3-hydroxybutan-2-one confirmed a high enzymatic activity of microorganisms [47,48]. These volatile compounds are mostly in line with those reported by different authors for dry-cured loin [47,49] and dry-cured ham [30,50,51].
Summarizing, the water, fat, ash, and salt content, as well as pH and water activity, of Pul and ZS dry-cured meat products changed depending on the breed and time of maturing. The count of lactic acid bacteria, as well as the contents of fatty acids, and basic amino acids (arginine, histidine, lysine), shaped the final pH. The high content of aliphatic and catechol amines in ZS products indicated more advanced proteolysis and decarboxylation of ornithine, lysine, and tyrosine, as well as a twofold greater share of heterocyclic amines via catabolism of histidine and tryptophan. The Pul and ZS products differed in the proportion of decanoic, lauric, stearic, and arachidic fatty acids. There were no significant differences in the sum of ω-6 acids, and the only acid slightly differing Pul and ZS products in the ω-3 group was CLA. The compounds that differentiated Pul and ZS products were obtained as a result of microbial activity and lipid oxidation. Volatile compounds shaped a buttery-fruity aroma with caramellic nuances and with the intensity characteristic for native pork meat.

4. Conclusions

Ecological production is associated with the high quality of raw materials, additives, health-promoting properties, safety, animal welfare, and sustainable development of agricultural production. The use of raw materials obtained from native breeds is becoming popular, which is associated with cultural heritage and supporting local producers. Based on these results, it can be mentioned that the meat of both native breeds can become the basis of dry-cured meats. The produced hams are of high quality, although the maturing processes in the assessment of many properties differ significantly. The meat of Pulawska and Zlotnicka Spotted breeds is an interesting target for native raw material studies with nonaccelerated metabolism of animal tissues.

Author Contributions

Conceptualization, E.W.; methodology, E.W.; investigation, J.S.-Z., M.P., K.N.-W., R.G., K.W., H.P., I.D.; resources, W.M.; writing—original draft preparation, E.W.; writing—review and editing, E.W.; visualization, E.W.; project administration, W.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financed by the Ministry of Science and Higher Education of the Republic of Poland.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available on request from the corresponding author.

Acknowledgments

This work was supported by the National Center for Research and Development (BIOSTRATEG2/297267/14/NCBR/2016), as well as by the Ministry of Science and Higher Education of the Republic of Poland.

Conflicts of Interest

Authors declare no conflict of interest.

References

  1. Babicz, M.; Blicharski, T.; Bajda, Z. (Eds.) Pigs of the Pulawy Breed, Heritage of the History of Polish Agriculture, 1st ed.; IGHZ PAN: Lublin, Poland, 2011. [Google Scholar]
  2. Babicz, M.; Kropiwiec, K.; Kasprzak, K.; Skrzypczak, E.; Hałabis, M. Analysis of the quality pork obtained from carcasses fatteners of Polish Landrace and Pulawska breed. Ann. UMCS Sect. EE 2013, 31, 1–7. [Google Scholar]
  3. Babicz, M.; Hałabis, M.; Skałecki, P.; Domaradzki, P.; Litwińczuk, A.; Kropiwiec-Domańska, K.; Łukasik, M. Breeding and performance potential of Pulawska pigs—A review. Ann. Anim. Sci. 2020, 20, 343–354. [Google Scholar] [CrossRef]
  4. Krupiński, J. Polish preserving breeds. In Atlas of Farm Animals Covered by the Protection Program in Poland, 1st ed.; IZ PIB: Kraków, Poland, 2012. [Google Scholar]
  5. Szyndler-Nędza, M.; Luciński, P.; Skrzypczak, E.; Szulc, K.; Bajda, Z. Pigs of native breeds—status of breeding and evaluation results. In Protection of Genetic Resources of Domestic Breeds of Pigs—Status of Breeding and Results of Evaluation for 2008, 1st ed.; Eckert, R., Ed.; IŻ PIB: Kraków, Poland, 2009. [Google Scholar]
  6. Szyndler-Nędza, M.; Luciński, P.; Szulc, K.; Bajda, Z. Pigs of native breeds—status of breeding and evaluation results. In Protection of Genetic Resources of Domestic Breeds of Pigs—Status of Breeding and Results of Evaluation for 2011, 1st ed.; Eckert, R., Ed.; IŻ PIB: Kraków, Poland, 2012. [Google Scholar]
  7. Kapelański, W.; Buczyński, T.J.; Bocian, M. Slaughter value and meat quality in the Polish native Zlotnicka Spotted pig. Anim. Sci. Pap. Rep. 2006, 24, 7–13. [Google Scholar]
  8. Grześkowiak, E.; Borys, A.; Borzuta, K.; Buczyński, J.; Lisiak, D. Slaughter value, meat quality and backfat fatty acid profile in Zlotnicka Spotted fatteners. Anim. Sci. Pap. Rep. 2009, 27, 115–125. [Google Scholar]
  9. Debrecéni, O.; Lípová, P.; Bučko, O.; Cebulska, A.; Kapelánski, W. Effect of pig genotypes from Slovak and Polish breeds on meat quality. Arch. Anim. Breed. 2018, 61, 99–107. [Google Scholar] [CrossRef] [Green Version]
  10. Szulc, K.; Skrzypczak, E.; Buczyński, J.T.; Stanisławski, D.; Jankowska-Małkosa, A.; Knecht, D. Evaluation of fattening and slaughter performance and determination of meat quality in Zlotnicka Spotted pigs and their crosses with the Duroc breed. Czech J. Anim. Sci. 2012, 57, 95–107. [Google Scholar] [CrossRef] [Green Version]
  11. Fabri, R.; Bergonzini, E. Performance of pure and crossbred spotted Poland pigs. Riv. Di Suinic. 1981, 22, 25–33. [Google Scholar]
  12. Jankowiak, H.; Kapelański, W.; Kwiatkowska, B.E.; Biegniewska, M.; Cebulska, A. Carcass and meat quality of Zlotnicka Spotted pigs in comparison to Polish Large White x Polish Landrace crossbred pigs. Res. Pig. Breed. 2009, 3, 4–6. [Google Scholar]
  13. Janiszewski, P.; Grześkowiak, E.; Lisiak, D.; Szulc, K.; Borzuta, K. Quality and technological suitability of meat from Złotnicka Spotted breed and its crossbreeds with the Duroc or Polish Large White. Sci. Ann. Pol. Soc. Anim. Prod. 2015, 11, 83–93. [Google Scholar]
  14. Cebulska, A. Quality and dietary value of pork meat of the Pulawska and Zlotnicka Spotted breeds, and commercial fattening pigs. Ann. Anim. Sci. 2018, 18, 281–291. [Google Scholar] [CrossRef] [Green Version]
  15. Migdał, W.; Kulig, R.; Walczycka, M.; Węsierska, E.; Zając, M.; Tkaczewska, J.; Kulawik, P.; Migdał, Ł.; Migdał, A.; Krępa-Stefanik, K. The quality of traditionally smoked tenderloins obtained from meat of native pig breeds. Biotechnol. Anim. Husb. 2020, 36, 49–61. [Google Scholar] [CrossRef] [Green Version]
  16. Skałecki, P.; Babicz, M.; Domaradzki, P.; Litwińczuk, A.; Hałabis, M.; Ruda, B. Basic chemical composition, colour and content of PAHs and nitrates in smoked pork products made from Puławska breed. Vet. Med. 2019, 75, 422–425. [Google Scholar] [CrossRef]
  17. Pugliese, C.; Sirtori, F. Quality of meat and meat products produced from southern European pig breeds. Meat Sci. 2012, 90, 511–518. [Google Scholar] [CrossRef]
  18. ISO 1442:2000 Meat and Meat Products—Determination of Moisture Content (Reference Method); Polish Committee for Standardization: Warszawa, Poland, 2000.
  19. PN-75-A-04018:1975/Az3:2002 Agri-Food Products—Determination of Nitrogen Content by the Kjeldahl Method and the Conversion of Protein; Polish Committee for Standardization: Warszawa, Poland, 2002.
  20. ISO 1444:2000 Meat and Meat Products—Determination of Fat-Free; Polish Committee for Standardization: Warszawa, Poland, 2000.
  21. ISO 1841-1:2002 Meat and Meat Products—Determination of Chloride Content—Part 1: Volhard Method; Polish Committee for Standardization: Warszawa, Poland, 2002.
  22. ISO 4833:2004/Ap1:2005 Microbiology of Food and Animal Feeding Stuffs—Horizontal Method for the Enumeration of Microorganisms; Colony-count technique at 30 °C; Polish Committee for Standardization: Warszawa, Poland, 2005.
  23. ISO 15214:2002 Microbiology of Food and Animal Feeding Stuffs—Horizontal Method for the Enumeration of Mesophilic Lactic Acid bacteria; Colony-count technique at 30 °C; Polish Committee for Standardization: Warszawa, Poland, 2002.
  24. ISO 6888:2001/A1:2004 Microbiology of Food and Animal Feeding Stuffs—Horizontal Method for the Enumeration of Coagulase-Positive Staphylococci (Staphylococcus Aureus and other Species); Part 1: Method using Baird-Parker agar; Polish Committee for Standardization: Warszawa, Poland, 2004.
  25. Kuchroo, C.V.; Ramilly, I.P.; Fox, P.F. Assessment of proteolysis in cheese by reaction with trinitrobenzensulphonic acid. J. Food Technol. 1983, 7, 129–133. [Google Scholar]
  26. Folch, J.; Lees, M.; Sloane, S.G.H. A simple method for the isolation and purification of total lipids from animal tissues. J. Biol. Chem. 1957, 226, 497–509. [Google Scholar] [CrossRef]
  27. Rosmini, M.R.; Perlo, F.; Pérez-Alvarez, J.A.; Pagán-Moreno, M.J.; Gago-Gago, A.; López-Santovena, F.; Aranda-Catalá, V. TBA test by an extractive method applied to ‘pate’. Meat Sci. 1996, 42, 103–110. [Google Scholar] [CrossRef]
  28. Calik, J.; Krawczyk, J.; Świątkiewicz, S.; Gąsior, R.; Wojtycza, K.; Połtowicz, K.; Obrzut, J.; Puchała, M. Comparison of the physicochemical and sensory characteristics of Rhode Island Red (R-11) capons and cockerels. Ann. Anim. Sci. 2017, 17, 903–917. [Google Scholar] [CrossRef] [Green Version]
  29. Olivares, A.; Navarro, J.L.; Salvador, A.; Flores, M. Sensory acceptability of slow fermented sausages based on fat content and ripening time. Meat Sci. 2010, 86, 251–257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Marušić, N.; Petrović, M.; Vidaček, S.; Petrak, T.; Medić, H. Characterization of traditional Istrian dry-cured ham by means of physical and chemical analyses and volatile compounds. Meat Sci. 2011, 88, 786–790. [Google Scholar] [CrossRef]
  31. Degnes, K.F.; Kvitvang, H.F.N.; Haslene-Hox, H.; Aasen, I.M. Changes in the profiles of metabolites originating from protein degradation during ripening of dry cured ham. Food Bioprocess Technol. 2017, 10, 1122–1130. [Google Scholar] [CrossRef]
  32. Spaziani, M.; Del Torre, M.; Stecchini, M.L. Changes of physicochemical, microbiological and textural properties during ripening of Italian low-acid sausages. Proteolysis, sensory and volatile profiles. Meat Sci. 2009, 81, 77–85. [Google Scholar] [CrossRef]
  33. Latorre-Moratalla, M.L.; Bosch-Fusté, J.; Bover-Cid, S.; Aymerich, T.; Vidal-Carou, M.C. Contribution of enterococci to the volatile profile of slightly-fermented sausages. LWT-Food Sci. Technol. 2011, 44, 145–152. [Google Scholar] [CrossRef]
  34. Ravyts, F.; De Vuyst, L.; Leroy, F. Bacterial diversity and functionalities in food fermentations. Eng. Life Sci. 2012, 12, 356–357. [Google Scholar] [CrossRef]
  35. Berardo, A.; Devreese, B.; De Maere, H.; Stavropoulou, D.A.; Van Royen, G.; Leroy, F.; De Smet, S. Actin proteolysis during ripening of dry fermented sausages at different pH values. Food Chem. 2017, 221, 1322–1332. [Google Scholar] [CrossRef] [Green Version]
  36. Węsierska, E.; Korzekwa, K.; Foks, S.; Mickowska, B. Influence of microflora composition on safety and colour parameters of „kumpia wieprzowa” during ripening. Pol. J. Vet. Sci. 2013, 16, 299–305. [Google Scholar] [CrossRef] [PubMed]
  37. Węsierska, E.; Szołtysik, M.; Rak, L. Physico-chemical, biochemical and microbiological properties of traditional Polish pork fermented products during ripening. Food Bioprocess Technol. 2013, 6, 2986–2995. [Google Scholar] [CrossRef]
  38. Mati, M.; Staruch, L.; Šoral, M. Use of NMR spectroscopy in the analysis of carnosine and free amino acids in fermented sausages during ripening. Chem. Pap. 2015, 69, 1319–1324. [Google Scholar] [CrossRef]
  39. Mora, L.; Gallego, M.; Escudero, E.; Reig, M.; Aristoy, M.C.; Toldrá, F. Small peptides hydrolysis in dry-cured meats. Int. J. Food Microbiol. 2015, 212, 9–15. [Google Scholar] [CrossRef] [Green Version]
  40. Zhou, H.Y.; Wang, Y.; Pan, D.D.; Cao, J.X.; Chen, Y.J.; Liu, Y.; Sun, Y.Y.; Ou, C.R. The changes in the proteolysis activity and the accumulation of free amino acids during Chinese traditional dry-cured loins processing. Food Sci. Biotechnol. 2017, 26, 679–687. [Google Scholar] [CrossRef]
  41. Innocente, N.; Biasutti, M.; Padovese, M.; Moret, S. Determination of biogenic amines in cheese using HPLC technique and direct derivation of acid extract. Food Chem. 2007, 101, 1285–1289. [Google Scholar] [CrossRef]
  42. Landeta, G.; de las Rivas, B.; Carrascosa, A.V.; Muñoz, R. Screening of biogenic amine production by coagulase-negative staphylococci isolated during industrial Spanish dry-cured ham processes. Meat Sci. 2007, 77, 556–561. [Google Scholar] [CrossRef] [Green Version]
  43. Lorenzo, J.M.; Martínez, S.; Franco, I.; Carballo, J. Biogenic amine content during the manufacture of dry-cured lacón, a Spanish traditional meat product: Effect of some additives. Meat Sci. 2007, 77, 287–293. [Google Scholar] [CrossRef]
  44. Martuscelli, M.; Pittia, P.; Casamassima, L.M.; Manetta, A.C.; Lupieri, L.; Neri, L. Effect of intensity of smoking treatment on the free amino acids and biogenic amines occurrence in dry cured ham. Food Chem. 2009, 116, 955–962. [Google Scholar] [CrossRef]
  45. Stadnik, J.; Dolatowski, Z. Biogenic amines content during extended ageing of dry-cured pork loins inoculated with probiotics. Meat Sci. 2012, 91, 374–377. [Google Scholar] [CrossRef] [PubMed]
  46. Virgili, R.; Saccani, G.; Gabba, L.; Tanzi, E.; Soresi Bordini, C. Changes of free amino acids and biogenic amines during extended ageing of Italian dry-cured ham. LWT—Food Sci. Technol. 2007, 40, 871–878. [Google Scholar] [CrossRef]
  47. Martin, D.; Antequera, T.; Muriel, E.; Perez-Palacios, T.; Ruiz, J. Volatile compounds of fresh and dry-cured loin as affected by dietary conjugated linoleic acid and monounsaturated fatty acids. Meat Sci. 2009, 81, 549–556. [Google Scholar] [CrossRef]
  48. Resconi, V.C.; Campo, M.M.; Montossi, F.; Ferreira, V.; Sañudo, C.; Escudero, A. Relationship between odor-active compounds and flavor perception in meat from lambs fed different diets. Meat Sci. 2010, 85, 700–706. [Google Scholar] [CrossRef] [PubMed]
  49. Ventanas, S.; Ventanas, J.; Estevéz, M.; Ruiz, J. Analysis of volatile molecules in Iberian dry-cured loins as affected by genetic, feeding systems and ingredients. Eur. Food Res. Technol. 2010, 231, 225–235. [Google Scholar] [CrossRef]
  50. Soto, E.; Hoz, L.; Ordóñez, J.A.; Hierro, E.; Herranz, B.; López-Bote, C.; Cambero, M.I. Impact of feeding and rearing systems of Iberian pigs on volatile profile and sensory characteristics of dry-cured loin. Meat Sci. 2008, 79, 666–676. [Google Scholar] [CrossRef]
  51. Théron, L.; Tournayre, P.; Kondjoyan, N.; Abouelkaram, S.; Santé-Lhoutellier, V.; Berdagué, J.L. Analysis of the volatile profile and identification of odour-active compounds in Bayonne ham. Meat Sci. 2010, 85, 453–460. [Google Scholar] [CrossRef]
Figure 1. The microbiological characteristics during ripening (mean, standard deviation, n = 3) of Pul (a) and ZS (b) products.
Figure 1. The microbiological characteristics during ripening (mean, standard deviation, n = 3) of Pul (a) and ZS (b) products.
Foods 10 01597 g001
Table 1. The least-squares means from two-factor analysis of variance of the basic chemical composition and physicochemical characteristics in relation to the type of dry-cured ham and the ripening period.
Table 1. The least-squares means from two-factor analysis of variance of the basic chemical composition and physicochemical characteristics in relation to the type of dry-cured ham and the ripening period.
CharacteristicsProduct (P)Ripening Period (Week) (RP)Interaction
PulZS024Source of Variance
MMMMMPRPP × RP
Water (%)72.15a71.02b74.29a72.19b69.12c***
Protein (%)18.83a20.63b18.19a18.79a20.85b**ns
Fat (%)4.99a4.53b3.43a4.77b6.44c***
Ash (%)3.02a4.04b1.31a3.96b4.55c***
Salt (%)1.89a1.70b0.42a2.47b2.64c***
pH5.52a5.69b5.69a5.68a5.33b***
aw0.97a0.96b0.98a0.96b0.95c***
Pul—Pulawska; ZS—Zlotnicka Spotted; M—mean; * F values of interaction significant at p < 0.05, n = 9; ns—not significant; a, b, c—values in rows with different letters differ significantly at p < 0.05, n = 9.
Table 2. One-factor analysis of the variance of the free amino-acid content of dry-cured ham.
Table 2. One-factor analysis of the variance of the free amino-acid content of dry-cured ham.
ProductPulZSProductPulZS
Amino Acid (mg/kg)MSDMSDAmino Acid (mg/kg)MSDMSD
Asparagine11.47a0.954.02b0.16Valine28.14a0.6922.69b0.89
Serine34.70a1.1332.68b1.28Methionine23.91a0.5918.81b0.73
Glutamine44.82a2.8223.18b0.91Lysine26.76a1.6722.92b0.90
Glycine23.08a1.2217.14b0.67Isoleucine25.63a0.7519.89b0.78
Histidine25.81a1.8920.65b0.81Leucine40.87a1.1532.83b1.28
Arginine30.04a1.8444.33b1.97Phenylalanine31.56a1.2234.65b1.35
Threonine43.66a1.5531.78b1.24Free amine groups (g/100 g)
Alanine47.45a1.1730.45b1.19in H2O0.61a0.060.70b0.00
Proline30.87a1.0821.33b0.83in PTA0.06a0.000.06a0.01
Tyrosine33.56a2.5730.97b1.21
Pul—Pulawska; ZS—Zlotnicka Spotted; PTA—phosphotungstic acid; M—mean; SD—standard deviation; a, b—values in rows with different letters differ significantly at p < 0.05, n = 9.
Table 3. One-factor analysis of the variance of the biogenic amine content of dry-cured ham.
Table 3. One-factor analysis of the variance of the biogenic amine content of dry-cured ham.
ProductPulZS
Biogenic Amines (mg/kg)MSDMSD
Tryptamine4.73a1.289.08b0.56
Phenylethylamine1.28a0.860.25a0.01
Putrescine3.97a2.1123.66b0.90
Cadaverine28.53a2.9454.31b3.03
Histamine87.89a16.70163.19b4.90
Tyramine11.07a0.3957.16b2.17
Spermine2.75a0.191.94b0.07
Spermidine21.69a1.0410.47b0.43
Pul—Pulawska; ZS—Zlotnicka Spotted; M—mean; SD—standard deviation; a, b—values in rows with different letters differ significantly at p < 0.05, n = 9.
Table 4. One-factor analysis of the variance of the fatty-acid content and TBA index of dry-cured ham.
Table 4. One-factor analysis of the variance of the fatty-acid content and TBA index of dry-cured ham.
ProductPulZSProductPulZS
Fatty Acids (%)MSDMSDFatty Acids (%)MSDMSD
10;00.12a0.010.08b0.0018; 2n-68.28a5.387.80a0.02
12;00.09a0.000.06b0.0018; 3n-60.05a0.010.06a0.00
14;01.24a0.201.04a0.0120; 3n-60.19a0.080.16a0.00
16;023.14a1.9525.91a0.2420; 4n-61.01a0.381.23a0.01
17;00.14a0.020.12a0.0122; 4n-60.18a0.040.16a0.00
18;010.09a0.3411.19b0.1322; 5n-60.02a0.000.01a0.00
20;00.09a0.000.13b0.01
Σ SFA34.91 38.53 Σ ω-69.73 9.42
14;10.02a0.010.02a0.0018; 3n-30.44a0.330.37a0.00
16;1n-74.04a0.923.33a0.0220; 4n-30.07a0.050.06a0.00
16;1n-90.73a0.010.71a0.0320; 5n-30.04a0.020.06a0.00
17;10.19a0.030.11b0.0122; 5n-30.16a0.090.16a0.01
18;1n-74.46a0.853.93a0.0822; 6n-30.08a0.030.07a0.00
18;1n-944.06a2.9942.22a0.28CLA0.06a0.000.05b0.00
20;10.58a0.020.61a0.03Σ ω-30.85 0.77
Σ MUFA54.08 50.93 Σ PUFA10.58 10.19
TBA (mg/kg)0.32a0.040.52b0.13
Pul—Pulawska; ZS—Zlotnicka Spotted; M—mean; SD—standard deviation; a, b—values in rows with different letters differ significantly at p < 0.05, n = 9.
Table 5. Volatile compounds of dry-cured ham—odor impression according to Flavornet; The Good Scents Company; Resconi et al. [48].
Table 5. Volatile compounds of dry-cured ham—odor impression according to Flavornet; The Good Scents Company; Resconi et al. [48].
Volatile Compound (ng/g)Ret. Index ZB-5Ret. Index ZB-WaxRet. TimeProduct PulProduct ZSFormationOdor Impression
MSDMSD
Ethanol (S)<5009441.7229.23a0.7832.53b7.56Microb. act. (fermentation)Alcoholic
Acetone (S)<5008161.8410.53a2.569.82b10.12Microb. act. (fermentation)Solvent, ethereal, apple
Thiirane, methyl-5968852.3911.37a1.2827.21b1.98Spices (garlic)
Acetic acid (S)64114442.98199.47a9.22151.31b15.67Microb. act. (fermentation)Pungent, vinegar
3-Methylbutanal6589163.1110.55a0.709.15b1.73Microb. act. (aa catabolism)Rancid, raw ham-like
1-Hydroxypropan-2-one67912913.525.92a0.264.87b1.72Lipid oxidationCaramellic, burnt
Allyl methyl sulfide6979533.869.56a0.549.38a0.28Spices (garlic)Sulfuric, alliaceous
3-Hydroxybutan-2-one71912794.4515.96a6.134.81b0.98Microb. act. (fermentation)Buttery, milky, fatty
3-Methylbutan-1-ol73712144.9841.77a0.1941.55a1.46Microb. act. (aa catabolism)Pungent, whisky-like
Pyridine (S)75711815.5416.43a2.9710.09b0.69PyrolysisFish-like
Toluen (S)76710326.0918.40a1.3415.19b1.77PyrolysisSweet
Hexanal (S)80110757.8211.28a7.2442.54b1.63Lipid oxidationAldehydic, fatty, fruity
2-Cyclopenten-1-one842134310.127.21a2.685.89b0.02Lipid oxidationCaramelized sugar
2-Methylcyclopentanone848118110.5511.67a1.3111.10a0.69Lipid oxidationMeaty, roasted beef-like
Diallyl sulfide865115011.869.83a0.905.60b1.58Spices (garlic)Alliaceous, metallic
2-Furanmethanol870166012.20112.18a2.8953.21b5.40SpicesBurnt
p-Xylen874112912.6211.21a0.3011.38a3.15PyrolysisAromatic
Hexan-1-ol881135813.125.57a1.006.72b1.13SpicesFreshly mown grass
o-Xylene (S)891117314.1610.53a0.365.99b0.73PyrolysisGeranium
2-Cyclopenten-1-one, 2-methyl-905135715.1328.15a1.0219.66b2.68Lipid oxidationCaramelized sugar
Allyl methyl disulfide915126915.635.93a2.567.25b5.99Spices (garlic)Alliaceous
γ-Butyrolactone915159915.6521.77a1.8021.30a3.00Lipid oxidationCreamy, caramel
2(5H)-Furanone917175415.797.38a0.566.30b3.73Lipid oxidationButtery
α-Pinene (S)936101316.7517.92a0.2817.26a1.17Spices (pepper)Camphoraceous, pine
Camphene950105317.568.39a0.536.11b0.26SpicesPungent
2-Furanmethanol, 5-methyl-963172818.369.40a0.844.80b0.19SpicesBurnt
2-Cyclopenten-1-one, 3-methyl-969150418.6315.48a1.758.70b0.27Lipid oxidationCaramelized sugar
β-Pinene (S)976108719.1714.89a0.4414.42a1.39Spices (pepper)Piney, eucalyptus
Oct-1-en-3-ol (S)985145319.656.10a0.627.76b0.08SpicesMushroom
β-Myrcene993116120.2012.53a0.649.42b0.39SpicesWoody, rosy, celery
α-Phellandrene1001115220.739.31a1.307.77b0.34Spices (pepper)Citrus, terpenic
3-Carene1008113421.0258.78a18.7147.89b8.01SpicesSweet, pungent
p-(o- or m-) Cymene1026126121.7913.22a0.1510.29b1.32SpicesTerpenic, citrus
Limonene (S)1030118321.9880.92a0.7465.34b6.47Spices (pepper)Menthol, orange
2-Ethylhexanol1034149222.185.68a0.376.02b0.88SpicesFresh floral, fatty
2,3-Dimethyl-2-cyclopenten-1-one1040152622.4719.38a0.3910.78b1.29Lipid oxidationCaramelized sugar
Phenylacetalaldehyde1045161822.7220.47a3.338.80b1.05Lipid oxidationAlmond-like
γ-Terpinene1061123423.4814.59a1.0811.55b2.58Spices (pepper)Terpenic, tropical
Linalool oxide (S)1077143924.238.53a0.654.90b0.22SpicesOrange, rose, terpenic
Diallyl disulfide1081148724.38107.78a10.2767.41b1.11Spices (garlic)Green onion meaty
Guaiacol1089186624.8999.79a5.0466.44b3.93PyrolysisWhisky, roasted coffee
Linalool (S)1100154225.44222.12a1.98144.86b1.83SpicesOrange, rose, terpenic
2-Phenylethanol1114192326.006.33a1.804.02b0.32SpicesRose, floral
Camphor1146149827.2915.48a0.9310.03b0.58SpicesCamphoraceous
Creosol1193194829.3246.14a5.3527.56b1.76PyrolysisCoal tar
4-Ethyl guaiacol1281203432.6915.27a1.688.51b0.55PyrolysisWhisky, roasted coffee
Eugenol1364217635.4173.98a8.3443.39b1.27SpicesClove-like, bacon-like
Caryophyllene1429157237.2311.83a0.7111.09a3.22SpicesPepper-like, citrus
Pul—Pulawska; ZS—Zlotnicka Spotted; M—mean; SD—standard deviation; a, b—values in rows with different letters differ significantly at p < 0.05, n = 9; (S)—authentic standard of the compound.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Węsierska, E.; Sobolewska-Zielińska, J.; Pasternak, M.; Niemczyńska-Wróbel, K.; Gąsior, R.; Wojtycza, K.; Pustkowiak, H.; Duda, I.; Migdał, W. Biochemical Properties Affecting the Nutritional Quality, Safety, and Aroma of Dry-Cured Products Manufactured from Meat of Rare Native Pig Breeds. Foods 2021, 10, 1597. https://doi.org/10.3390/foods10071597

AMA Style

Węsierska E, Sobolewska-Zielińska J, Pasternak M, Niemczyńska-Wróbel K, Gąsior R, Wojtycza K, Pustkowiak H, Duda I, Migdał W. Biochemical Properties Affecting the Nutritional Quality, Safety, and Aroma of Dry-Cured Products Manufactured from Meat of Rare Native Pig Breeds. Foods. 2021; 10(7):1597. https://doi.org/10.3390/foods10071597

Chicago/Turabian Style

Węsierska, Ewelina, Joanna Sobolewska-Zielińska, Małgorzata Pasternak, Katarzyna Niemczyńska-Wróbel, Robert Gąsior, Krzysztof Wojtycza, Henryk Pustkowiak, Iwona Duda, and Władysław Migdał. 2021. "Biochemical Properties Affecting the Nutritional Quality, Safety, and Aroma of Dry-Cured Products Manufactured from Meat of Rare Native Pig Breeds" Foods 10, no. 7: 1597. https://doi.org/10.3390/foods10071597

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

Article Metrics

Back to TopTop