Explore
Communities in English
Advertise on Engormix

Characteristics of Nutrition and Metabolism in Dogs and Cats

Published: April 26, 2024
By: Peng Li 1 and Guoyao Wu 2 / 1 North American Renderers Association, Alexandria, VA 22314, USA; 2 Department of Animal Science, Texas A&M University, College Station, TX 77843, USA.
Summary

Domestic dogs and cats have evolved differentially in some aspects of nutrition, metabolism, chemical sensing, and feeding behavior. The dogs have adapted to omnivorous diets containing taurine-abundant meat and starch-rich plant ingredients. By contrast, domestic cats must consume animal-sourced foods for survival, growth, and development. Both dogs and cats synthesize vitamin C and many amino acids (AAs, such as alanine, asparagine, aspartate, glutamate, glutamine, glycine, proline, and serine), but have a limited ability to form de novo arginine and vitamin D3. Compared with dogs, cats have greater endogenous nitrogen losses and higher dietary requirements for AAs (particularly arginine, taurine, and tyrosine), B-complex vitamins (niacin, thiamin, folate, and biotin), and choline; exhibit greater rates of gluconeogenesis; are less sensitive to AA imbalances and antagonism; are more capable of concentrating urine through renal reabsorption of water; and cannot tolerate high levels of dietary starch due to limited pancreatic α-amylase activity. In addition, dogs can form sufficient taurine from cysteine (for most breeds); arachidonic acid from linoleic acid; eicosapentaenoic acid and docosahexaenoic acid from α-linolenic acid; all-trans-retinol from β-carotene; and niacin from tryptophan. These synthetic pathways, however, are either absent or limited in all cats due to (a) no or low activities of key enzymes (including pyrroline-5-carboxylate synthase, cysteine dioxygenase, ∆6-desaturase, β-carotene dioxygenase, and quinolinate phosphoribosyltransferase) and (b) diversion of intermediates to other metabolic pathways. Dogs can thrive on one large meal daily, select high-fat over low-fat diets, and consume sweet substances. By contrast, cats eat more frequently during light and dark periods, select high-protein over low-protein diets, refuse dry food, enjoy a consistent diet, and cannot taste sweetness. This knowledge guides the feeding and care of dogs and cats, as well as the manufacturing of their foods. As abundant sources of essential nutrients, animal-derived foodstuffs play important roles in optimizing the growth, development, and health of the companion animals.

Keywords Dogs · Cats · Nutrition · Metabolism · Feeding · Health · Animal-sourced foodstuffs

4.1 Introduction

The metabolism of most nutrients in domestic dogs and cats is similar to that in other mammals (Baker and Czarnecki-Maulden 1991). Thus, the qualitative dietary requirements of dogs and cats for most nutrients [e.g., amino acids (AAs) that are not formed de novo in animal cells] are similar to those for omnivores (e.g., humans and pigs). However, dogs and cats have a relatively short digestive tract (He et al. 2024) and have evolved to have some unique feeding behaviors and metabolic characteristics that are distinct from most of other nonruminant mammals such as pigs, rats, and humans (Legrand-Defretin 1994; Morris 2002). In addition, dogs differ from cats in some of these physiological and nutritional aspects (Che et al. 2021; Oberbauer and Larsen 2021). For example, the dog has adapted to omnivorous diets containing taurine-rich meat and starch-rich plant ingredients, but the cat must consume at least a portion of animal-sourced foods or the same essential nutrients from synthetic supplements for survival, growth, and development because of their limited ability to synthesize taurine (Baker and Czarnecki-Maulden 1991). In addition, there are peculiar differences in the syntheses and requirements of certain nutrients between dogs and cats, including taurine (essential for tissue integrity; Pion et al. 1987), arginine (essential for maintaining the urea cycle in an active state; Wu and Morris 1998), ω3 and ω6 polyunsaturated fatty acids (PUFAs; essential for cell structure and function; Bauer 2007), vitamin A (essential for retinal health; Case et al. 2011), and niacin (essential for nutrient metabolism; NRC 2006). Therefore, the concentrations of many AAs, fatty acids, and their metabolites in plasma differ between dogs and cats fed the same diet (Hall 2018b). Regardless of their sources of foods, adequate knowledge of nutrient metabolism and requirements by dogs and cats is crucial to ensure their optimal growth, development, and health. The major objective of this article is to integrate the metabolic characteristics of dogs and cats with their nutrition and feeding management.
4.2 Energy Metabolism
Energy is defined as the capacity to do work (Wu 2018). Energy metabolism is the sum of the metabolism of carbohydrates, lipids, and AAs in animals. Resting metabolic rate (measured after an overnight fast) of a dog and a cat, like other animals, is an estimate of its minimum energy requirement. In dogs living in a thermoneutral environment with a 12-h light period/a 12-h dark period, their heat production rate varies with a circadian rhythm, with the lowest rate in the morning (e.g., 124 kcal/kg BW0.75/day in beagles), gradually increasing to a maximum in the evening (e.g., 195 kcal/kg BW0.75/day in beagles), and then declining through the night (Besch and Woods 1977). Such changes in heat production are largely independent of food intake. The rate of heat production by dogs during the dark period is similar to that during the light period (Woods and Besch 1971). In contrast to dogs, no circadian rhythm in heat production was detected in cats (Riond et al. 2003). Furthermore, thermoneutral zones differ between dogs and cats, and even among different breeds of the same species or different sizes of the same breed (Hill and Scott 2004; Middleton et al. 2017). At an ambient temperature of 25 °C, dogs, but not cats, are within their thermoneutral zone.
In dogs and cats, as in other mammals, the maintenance energy requirement supports energy equilibrium over a long period of time and is calculated as 70.5 × BW0.73kcal (BW is body weight in kg; Brody et al. 1934), which can be simplified as 70 × BW0.75kcal (BW is body weight in kg; Kleiber 1961). Metabolizable energy (ME) is commonly used to express the requirements of dogs and cats for dietary energy. The energy requirements of adult dogs and cats may be related to their body surface areas [0.105 and 0.110 m2/kg of BW0.67, respectively (Hill and Scott 2004)]. NRC (2006) recommended that intact active adult dogs, normal adult lean cats, and obese cats be fed ∼130 kcal ME/kg BW0.75/day, 100 kcal ME/kg BW0.67/day, and 130 kcal ME/kg BW0.4/day, respectively. Through an extensive meta-analysis, Bermingham et al. (2010) indicated that the maintenance energy requirement of adult cats based on their BW alone may not be accurate and many factors (e.g., age, sex, neuter status, and composition) other than BW should be taken into consideration. Their proposed allometric equations for predicting the maintenance energy requirements of adult light, normal, and heavy cats were 53.7 kcal/kg BW–1.061, 46.8 kcal/ kg BW–1.115, and 131.8 kcal/kg BW–0.366, respectively, and were 58.4 kcal/kg lean mass−1.140.
Chronic imbalances between energy intake and expenditure result in obesity in dogs and cats, as in other animals (Larsen and Oberbauer 2023; Zoran 2023). As reported for many other mammals (e.g., humans, rats, and sheep), brown adipose tissue (BAT) is present in neonatal dogs and nearly absent from adult dogs (Holloway et al. 1985). Likewise, there is more BAT in neonatal cats than in adult cats (Clark et al. 2013; Loncar and Afzelius 1989). Maintaining BAT in an active state plays an important role in preventing obesity in cats. A recent study has indicated the browning and beiging of adipose tissue in dogs (Lyer et al. 2019). Both shivering and non-shivering mechanisms are responsible for increasing heat production in dogs exposed to a cold environment. Epinephrine and norepinephrine stimulate lipolysis in both BAT and white adipose tissues to provide fatty acids and glycerol as metabolic fuels (Wu 2018).

4.3 Lipid Metabolism in Dogs and Cats

4.3.1 Overview of Lipid Metabolism

Lipids are hydrocarbon compounds (e.g., fats, long-chain fatty acids, and cholesterol) that are soluble in organic solvents and, except for short- and medium-chain fatty acids, are insoluble in water. Fats contribute to the palatability and texture of pet foods, and also have important metabolic functions in animals (Wu 2018). The carbon of a fatty acid can be counted from either the carboxyl group (the ∆ nomenclature system) or the methyl group (the ω nomenclature system). In dogs and cats, as in other animals, no double bond can be introduced beyond the ∆9 carbon, and therefore, a ω3, ω6, ω7, or ω9 fatty acid (with the first double bond appearing in the 3rd, 6th, 7th, and 9th carbon counted from the methyl group, respectively) remains in the same class of the fatty acid despite its desaturation and elongation (Baker and Czarnecki-Maulden 1991). Animal-sourced, but not plant-sourced foodstuffs contain cholesterol. Cholesterol is an essential component of cell membranes and plays an important role in the metabolism, growth, development, and reproduction of animals, but an excessive amount of this lipid in the blood increases risk for cardiovascular disorders (Wu 2018).
Because no double bond can be formed beyond the ∆9 position in all animals, both α-linolenic acid (C18:3, ω3; ∆9,12,15) and linoleic acid (C18:2, ω6; ∆9,12) are not synthesized by dogs and cats (Table 4.1). These two fatty acids are integral components of the plasma membrane of cells (including erythrocytes and respiratory epithelial cells) for cell integrity, transport, and function (Bryan et al. 2001; Lemaitre et al. 2009), thereby exerting enormous physiological functions in tissues. In addition, α-linolenic acid and linoleic acid are the precursors for the synthesis of eicosapentaenoic acid (EPA; C20:5, ω3; ∆5,8,11,14,17) plus docosahexaenoic acid (DHA; C22:6, ω3; ∆4,7,10,13,16,19) and arachidonic acid (C20:4, ω6; ∆5,8,11,14), respectively (Wu 2018). EPA, DHA, and arachidonic acid are also essential components of cell membranes to affect cell integrity, signaling, and homeostasis. Thus, both linolenic acid and linoleic acid must be supplied in the diets for dogs and cats to ensure their survival, growth, development, and health (Table 4.2).
Beef tallow and butter contain low content of linoleic acid and α-linolenic acid. Therefore, when fed to dogs and cats, these two foodstuffs should be used along with other ingredients such as corn, soybean, and safflower oils that are rich in linoleic acid and α-linolenic acid (NRC 2006). As reported for humans (Eyjolfson et al. 2004), conjugated linoleic acid (CLA) improves metabolic profiles in obese or overweight dogs (Schoenherr and Jewell 1999) and cats (Bartges and Cook 1999). For example, dietary supplementation with 0.3% CLA to overweight dogs for 120 days decreased fat deposition in subcutaneous tissue and the mass of whole-body white fat mass but increased blood levels of high-density lipoproteins (Rivera et al. 2011). CLA-rich foods are ruminant meat, fats, and dairy products, with the predominant CLA isomer (> 90%) being c9,t11 (Larsen et al. 2003).

4.3.2 Lipid Metabolism in Dogs

Like most other mammals (e.g., humans, pigs, and rats), dogs can introduce one or more double bonds to a long-chain fatty acid between its ∆1 and ∆9 carbons through the action of ∆-desaturases (enzymes in the smooth endoplasmic reticulum). Dogs have ∆6 desaturase activity and, therefore, can readily convert linoleic acid into arachidonic acid (Dunbar and Bauer 2002). Likewise, growing and young adult dogs can actively form EPA and DHA from α-linolenic acid. However, the synthesis of EPA and DHA is insufficient to meet the needs of these dogs (NRC 2006). Furthermore, as reported for other animals (Bordoni et al. 1988), ∆6 desaturase activity in the tissues of dogs may decline markedly with increasing age, thereby reducing EPA and DHA syntheses in older dogs. By contrast, dogs express a high activity of stearoyl-CoA desaturase (a ∆9 desaturase in the smooth endoplasmic reticulum) to readily convert palmitic acid (C16:0; a saturated long-chain fatty acid) into ω7 and ω9 unsaturated fatty acids (Bauer 2007). The desaturation and elongation of long-chain fatty acids to form mono- and polyunsaturated fatty acids in the smooth endoplasmic reticulum are shown in Fig. 4.1.
Dogs have a wide tolerance for dietary lipids (e.g., at least 40% fat in their diets) and can be maintained on diets containing 5–8% lipids on the DM basis (NRC 2006). A minimum content of 5% lipids in a canine diet is adequate if it provides sufficient essential fatty acids. To meet their requirement for energy, sled dogs are often fed a diet containing at least 50% fat during the racing season (Hill 1998; Loftus et al. 2014). During fasting, dogs mobilize fat in white adipose tissue and rapidly utilize ketone bodies as metabolic fuels (De Bruijne and van den Brom 1986; McCue 2010). This may explain, in part, why the concentrations of acetoacetate and β-hydroxybutyrate in the plasma of fasting dogs are low (< 0.15 mM) during brief and long-term fasting (De Bruijne and van den Brom 1986).
Table 4.1 De novo formation of nutrients in dogs, cats, and pigs
Table 4.1 De novo formation of nutrients in dogs, cats, and pigs
Table 4.2 Dietary requirements of nutrients by dogs, cats, and pigs
Table 4.2 Dietary requirements of nutrients by dogs, cats, and pigs
Fig. 4.1 Elongation of the unsaturated fatty acid chain beyond C18, as well as the formation of monosaturated fatty acids in the smooth endoplasmic reticulum. This metabolic pathway also requires the addition of acetyl-CoA as malonyl-CoA to the hydrocarbon chain, as well as NADPH. The shortening of a polyunsaturated fatty acid chain from C24 to C22 through β-oxidation in the peroxisome is also shown. All desaturases are localized in the smooth endoplasmic reticulum. ∆5-D/ E = ∆5-desaturase and elongase; EL = elongase; SCD = stearoyl-CoA desaturase (∆9 desaturase)
Fig. 4.1 Elongation of the unsaturated fatty acid chain beyond C18, as well as the formation of monosaturated fatty acids in the smooth endoplasmic reticulum. This metabolic pathway also requires the addition of acetyl-CoA as malonyl-CoA to the hydrocarbon chain, as well as NADPH. The shortening of a polyunsaturated fatty acid chain from C24 to C22 through β-oxidation in the peroxisome is also shown. All desaturases are localized in the smooth endoplasmic reticulum. ∆5-D/ E = ∆5-desaturase and elongase; EL = elongase; SCD = stearoyl-CoA desaturase (∆9 desaturase)

4.3.3 Lipid Metabolism in Cats

In contrast to dogs, cats have a very limited ∆6 desaturase activity (Pawlosky et al. 1994; Rivers et al. 1975). Thus, when cats are fed a diet rich in linoleic acid, there is a modest increase in: (1) tissue concentrations of eicosadienoic acid (C20:2, ω6; ∆11,14) due to chain elongation, and (2) the formation of sciadonic acid (C20:3, ω6; ∆5,11,14) due to the action of ∆5 desaturase activity (Sinclair et al. 1981). These metabolic pathways are shown in Fig. 4.1. Sciadonic acid is a biomarker for limited ∆6 desaturase activity but active ∆5 desaturase in animal tissues (Bauer 1997). Accordingly, cats generate only a limited amount of arachidonic acid from linoleic acid (Sinclair et al. 1981), but dogs and omnivores (e.g., humans and pigs) readily convert linoleic acid to arachidonic acid as noted previously. Cats have a greater requirement for dietary arachidonic acid than dogs. In addition, a low activity of ∆6 desaturase in cats limits the conversion of α-linolenic acid into EPA and DHA. Thus, diets for cats must contain EPA and DHA. Daily supplementation of DHA and EPA is recommended for pregnant and lactating cats for the proper brain and nervous system development of their offspring (Vuorinen et al. 2020).

4.4 Carbohydrate Metabolism in Dogs and Cats

4.4.1 Overview of Carbohydrate Metabolism

Glucose is the exclusive energy source for the brain, red blood cells, kidney medulla, and retinal cells of dogs and cats in fed and post-absorptive states (Wu 2018). Because the brain is relatively large in newborn dogs and cats (accounting for 2.74% and 3.61% of BW, respectively; Latimer 1967), a large amount of glucose is used by the brain of puppies and kittens. For comparison, the brain of adult dogs and cats represents 0.85% and 0.80% of BW, respectively (Latimer 1967), which is greater than the brain of cattle and sheep (0.2% of BW, Wu 2018). In addition, glucose is a major metabolic fuel for immune cells (e.g., lymphocytes and macrophages), as well as the primary source of NADPH for anti-oxidative reactions and nitric-oxide synthesis. As noted previously, dogs, unlike cats, have adapted during evolution to omnivorous diets that contain both animal- and plant-sourced foods (Baker and Czarnecki-Maulden 1991). Gluconeogenesis (formation of glucose from non-glucose substrates such as AAs, glycerol, lactate, and odd-carbon number fatty acids) plays an important role in maintaining glucose homeostasis and carbon balances in all mammals, including both dogs and cats (Wu 2018). There are no established requirements of dogs and cats for dietary carbohydrates.

4.4.2 Carbohydrate Metabolism in Dogs

When diets do not provide sufficient starch, glycogen, or glucose, dogs must synthesize glucose from glucogenic AAs, lactate, and glycerol in their liver and kidneys (Belo et al. 1976). In the canine liver and kidneys, phosphoenolpyruvate carboxykinase (PEPCK) is localized to both the cytosol and mitochondria (e.g., equal distribution in the liver but 50% and 65%, respectively, in the kidney under fed conditions) (Croniger et al. 2002; Wolf and Mehlman 1972). Such an intracellular compartmentation of PEPCK allows for the use of all glucogenic substrates (including glucogenic AAs) for glucose synthesis (Feng et al. 1996; Wolf and Mehlman 1972). Note that if PEPCK is exclusively localized to the mitochondria of the liver and kidneys, there is no formation of glucose from AAs due to the absence of NADH provision, as reported for the liver of fed or fasted chickens (Watford 1985). Interestingly, the activities of mitochondrial PEPCK (converting oxaloacetate into phosphoenolpyruvate) and pyruvate carboxylase (converting pyruvate into oxaloacetate) in the liver and kidneys of dogs are enhanced by the feeding of a carbohydrate-free diet (Feng et al. 1996). Intakes of dietary carbohydrates (including starch and fiber) affect postprandial concentrations of glucose and insulin in the plasma of dogs (Carciofi et al. 2008).
In dogs, water-soluble fibers (e.g., plant pectin, fructans, gums, psyllium, and β-glucan) and water-insoluble fibers (e.g., plant cellulose and hemicellulose) play an important role in their intestinal health (Nogueira et al. 2019). Soluble fibers in oats, peas, beans, apples, citrus fruits, carrots, barley, and psyllium form a viscous solution upon contact with water and are highly fermentable in the large intestine (NRC 2006; Silvio et al. 2000). Insoluble fibers in plant cell walls, wheat bran, vegetables, and whole grains retain water, do not form a viscous solution, and are much less fermentable than soluble fibers in the gut. Insoluble fibers stimulate intestinal peristalsis, increase fecal mass, and decrease the transit time of food through the GIT (Schneeman 1994). Some fermentable fibers (e.g., fructooligosaccharides) are beneficial prebiotics for the intestinal health of dogs (Montserrat-Malagarriga et al. 2024; Nogueira et al. 2019).

4.4.3 Carbohydrate Metabolism in Cats

As noted previously, cats consume meat (containing low glycogen content, < 5% on a DM basis) and no starch. Cats, just like dogs, have little or no salivary α-amylase activity (McGeachin and Akin 1979). In addition, pancreatic α-amylase activity is low in cats as compared with dogs. For example, pancreatic α-amylase activity in adult cats is only 2.3% and 2% of that in adult dogs and pigs, respectively (Kienzle 1993a), and the abundance of glucose transporters in the small intestine of cats is much lesser than that in dogs (Batchelor et al. 2011). Therefore, cats have a much lower ability to digest and use dietary starch in the small intestine than dogs, and their consumption of carbohydrate-rich diets can result in intestinal inflammation (Morris et al. 1977; Verbrugghe and Hesta 2017).
Because of a limited intake of glycogen, starch, and sugars in their natural foods, cats primarily depend on constant gluconeogenesis (mainly from AAs) to provide glucose (Legrand-Defretin 1994). This metabolic pathway is particularly significant for newborn cats with relatively a large brain (accounting for 3.62% of BW; Latimer 1938), where a large amount of glucose is used via glycolysis and the Krebs cycle. For comparison, the brain of adult cats represents 0.80% of BW (Latimer 1967), which is similar to that for adult dogs. Unlike dogs, glucokinase activity is absent from the feline liver (Washizu et al. 1999) just like the ruminant liver (Wu 2018). By contrast, the activities of other glycolytic enzymes (e.g., hexokinase I, phosphofructokinase, and pyruvate kinase) in the feline liver are substantially greater than those in the canine liver (Washizu et al. 1999). Furthermore, the activities of rate-controlling enzymes for gluconeogenesis (e.g., pyruvate carboxylase, fructose1,6-bisphosphatase, and glucose-6-phosphatase) in the feline liver are also much greater than those in the canine liver (Washizu et al. 1999). This is consistent with a greater rate of whole-body glucose production in cats than in dogs (Legrand-Defretin 1994). Interestingly, hepatic PEPCK activity is not altered in cats that are fasted after consuming high-protein diets (Kettelhut et al. 1980) or after the protein content of their diet is increased from 17.5% (low) to 70% (high) (Rogers et al. 1977). These results support the view that cats express high basal activities of hepatic gluconeogenic enzymes to ensure the rapid conversion of excess dietary AAs into glucose. In cats, gluconeogenesis plays an essential role in the provision of glucose to their brain, red blood cells, and immunocytes and therefore their survival.
Just like dogs, carbohydrates (including water-soluble and insoluble fibers) are also extensively fermented by microbes in the large intestine of cats (Brosey et al. 2000; Legrand-Defretin 1994; Morris et al. 1977). For example, in adult cats, the prececal apparent digestibility of starch (e.g., cooked corn starch, 72%; raw corn starch, 46%; raw potato starch, 0%) is lower than its total apparent (fecal) digestibility (cooked corn starch was digested to nearly 100%, raw corn starch to 78%, raw potato starch to 36%), but the pattern of apparent digestibility of raw starch is similar to that for the cooked starch (Kienzle 1993b). This may help to explain the observation that the concentration of glucose in adult cats fed a 31.7%-starch diet was continuously elevated from 4.6 mM at 3 h after the meal, compared with the pre-meal baseline (4.8 mM) to 6.2 mM at 19 h after feeding, whereas the concentration of glucose in adult cats fed a 23%-starch diet did not differ from the pre-meal level or during the period of 3 and 19 h after the meal (Hewson-Hughes et al. 2011). Due to its microbial fermentation, starch led to an acidification of the large bowel chyme and feces, and the effect of dietary undigested starch can be alleviated by oral antibiotics in cats (Kienzle 1993b). This may explain, in part, why cats should not be fed dog food and require carefully formulated diets. Commercial diets for cats generally contain a small amount of water-soluble and insoluble fibers (a total of 3–5% fibers, DM basis) to maintain intestinal health (Zoran 2023).

4.5 Protein and AA Metabolism in Dogs and Cats

4.5.1 Overview of Protein and AA Metabolism

Both dogs and cats have a high ability to digest dietary protein in the gastrointestinal tract and degrade dietary AAs in a tissue-specific manner (Galim et al. 1980; Beliveau and Freedland RA 1982; Hammer et al. 1996; Hu et al. 2021; Li and Wu 2023a; Wu 1998). These mammals, just like all other animals, cannot form de novo the carbon skeletons of Cys, His, Ile, Leu, Lys, Met, Phe, Thr, Trp, Tyr, and Val, but can convert Phe into Tyr, Met into Cys, and Trp to serotonin (Wu 2022). Some metabolites of tyrosine (melanins) are crucial for the color of hair and skin health (Anderson et al. 2002; Biourge and Sergheraert 2002; Yu et al. 2001), whereas serotonin regulates animal behavior (Landsberg et al. 2017) and intestinal function (Chiocchetti et al. 2021). In addition, these two animal species, young or adult, are not capable of synthesizing sufficient arginine from glutamine/glutamate and proline (Burns et al. 1981; Czarnecki and Baker 1984; Weber et al. 1977; Yu et al. 1996). This is likely due to a deficiency of pyrroline-5-carboxylate synthase, proline oxidase, carbamoylphosphate synthetase I, and/or N-acetylglutamate synthetase as well as ornithine aminotransferase in enterocytes of the canine and feline small intestines (Dillion and Wu 2021; Morris 2002; Rogers and Phang 1985; and Wu and Morris 1998). In addition, most breeds of dogs,with possible exceptions of a few, can synthesize sufficient taurine when fed methionine- and cysteine-adequate diets (Jacobsen and Smith 1968). Interestingly, golden retrievers, American Cocker Spaniels, and a small population (1.3–2.5%) of the Newfoundlands may be genetically predisposed to taurine deficiency possibly due to gene mutations (Backus et al. 2003, 2006; Kaplan et al. 2018; Kittleson et al. 1997). By contrast, all breeds of cats have a very limited ability to convert cysteine into taurine due to low activities of cysteine dioxygenase and cysteinesulphinic acid decarboxylase in the liver (Knopf et al. 1978). Thus, Arg, His, Ile, Leu, Lys, Met, Phe, Thr, Trp, and Val must be included in canine and feline diets; taurine must be provided to all cats and some breeds of dogs. As reported for pigs, rats, fish and crustaceans (He and Wu 2022; He et al. 2023; Hou et al. 2015, 2016; Hou and Wu 2015; Li et al. 2020, 2021b,c, 2023; Wu 2009; Wu and Li 2022), dogs and cats may have dietary requirements for AAs (e.g., glutamate, glutamine, glycine, and proline) that are synthesized de novo and also use dietary AAs as major metabolic fuels. Their quantitative amounts have recently been suggested for both animal species during the growth period and in adulthood (Li and Wu 2023a). Growing evidence shows that AAs are signaling molecules to activate the mechanistic target of rapamycin pathway for initiating protein synthesis in animal cells (Rezaei and Wu 2022), as well as regulating energy sensing and metabolism (Jobgen et al. 2022a, b, 2023). Both dogs and cats have dietary requirements for AAs but not protein or nitrogen. In the following sections, we focus on arginine and cysteine in canine and feline nutrition, as well as AA imbalances and antagonisms, metabolic adaptation to AA intakes, and AA requirements.

4.5.2 Arginine Metabolism in Dogs and Cats

In most mammals, including dogs and cats, the small intestine synthesizes de novo and releases citrulline, which is taken up by extraintestinal tissues (primarily the kidneys and, to a lesser extent, all other tissues) for arginine synthesis (Levillain et al. 1996; Wu and Morris 1998). However, endogenous synthesis of arginine is limited in both animal species, as noted previously. All mammalian tissues contain arginase for the hydrolysis of arginine into urea and ornithine, with ornithine being either converted into proline, glutamate, polyamines, agmatine, and NO or oxidized to CO2 in a cell-specific manner (Wu et al. 2016b). Thus, the dietary intake, endogenous synthesis, and catabolism of arginine regulate its homeostasis in the body.
Consistent with the lower rate of arginine synthesis in cats than in dogs, the milk of cats contains more arginine than the milk of dogs (Davis et al. 1994a, b) to support feline survival, growth, and development. In addition, the content of arginine in the milk of dogs is greater than that in the milk of herbivores (e.g., cows) and omnivores (e.g., humans and pigs) to compensate for a lower rate of arginine synthesis in the dogs (Heinze et al. 2014). Current commercial vegan foods may not provide sufficient arginine for cats (Zafalon et al. 2020b).
A deficiency of arginine results in decreased food intake, impaired ureagenesis, hyperammonemia, abnormal hemodynamics, hypertension, severe emesis, frothing at the mouth, muscle tremors, and cataract in dogs (Burns et al. 1981; Czarnecki and Baker 1984; Ranz et al. 2002). Cats have a lower rate of endogenous arginine synthesis, ingest more AAs, and are more sensitive to a deficiency of dietary arginine than dogs. Thus, a lack of dietary arginine rapidly causes hyperammonemia (occurring within 1–3 h after feeding), vomiting, neurological signs, severe emesis, ataxia, tetanic spasms, and death in cats (Morris and Rogers 1978). Dietary supplementation with arginine, citrulline, or ornithine to dogs and cats can prevent hypoargininemia-induced hyperammonemia and mortality. Dietary or intravenous administration of arginine or citrulline, but not ornithine, can restore growth in young cats fed arginine-free or deficient diets (Czarnecki and Baker 1984). This is due to the complex compartmentalization of ornithine metabolism in enterocytes, as extracellular ornithine is preferentially channeled into the formation of proline but not citrulline in these cells (Wu and Morris 1998). Greater protein requirements by cats than dogs may be explained, in part, by a lower ability of cats to synthesize arginine than dogs and a need for the greater provision of arginine from feline than canine diets (Li and Wu 2023a).

4.5.3 Cysteine Metabolism in Dogs and Cats

All cats and some breeds of dogs have a limited ability to synthesize taurine from cysteine (a metabolite of methionine; Fig. 4.2), as noted previously. This metabolic pathway occurs in the liver of dogs and cats, with dietary cysteine being capable to replace up to 50% dietary methionine (Che et al. 2021; Oberbauer and Larsen 2021). In all animals, the synthesis of cysteine requires methionine and serine, as the sulfur atom of cysteine is derived from methionine, whereas serine provides both the amino group and the whole carbon skeleton of cysteine (Wu 2022). Of particular note, taurine is abundant in animals-sourced foodstuffs but is absent from plants (Hou et al. 2017, 2019; Li et al. 2021a). In dogs and cats, the primary bile acids are conjugated exclusively and almost exclusively with taurine, respectively, to form bile salts, which are required for the digestion and absorption of dietary lipids (including long-chain fatty acids and lipid-soluble vitamins). In healthy dogs, the primary bile acids are cholic acid, deoxycholic acid, and chenodeoxycholic acid; their tauroconjugated forms account for approximately 73%, 20%, and 6%, respectively, of the total bile acid pool in the gallbladder (Washizu et al. 1994); and there are no glycineconjugated bile acids in the liver or bile (Wildgrube et al. 1986). In healthy cats, the major primary bile acids are cholic acid and deoxycholic acid, and their distribution percentages (%) are 98.4% taurine-conjugated form (79.1% cholic acid and 19.3% deoxycholic acid), 1.1% glycine-conjugated cholic acid, and 0.5%of unconjugated cholic acid (Rabin et al. 1976).
Because the content of methionine and cysteine in plant-sourced ingredients is generally lower than that in animal-sourced ingredients, an inadequate intake of methionine and cysteine in plant-based foods can result in a deficiency of taurine in both dogs and cats (Cavanaugh et al. 2021; Knopf et al. 1978; McCauley et al. 2020). A deficiency of taurine in these companion animals can result in dilated cardiomyopathy that is characterized by thin heart muscle and enlarged chambers (Cavanaugh et al. 2021; McCauley et al. 2020; Morris 2002). Additional symptoms in cats include heart failure, central retinal degeneration, blindness, deafness, poor reproduction, and impaired immune responses (Morris 2002; Pion et al. 1992). All foods for cats and some breeds of dogs must provide sufficient taurine to maintain the integrity and health of their tissues. Morris et al. (1990) reported that the content of 0.12% taurine (on a DM basis) in commercial expanded (dry) cat foods was sufficient for cats, but the content of taurine in canned diets should be increased to 0.20–0.25% (DM basis) because heating during the canning process forms products that promotes the loss of taurine via the enterohepatic circulation.
Another metabolite of cysteine via cooperation of the liver and kidneys in both dogs and cats is isovalthine, the formation of which also requires leucine (Wu 2022). Interestingly, male cats produce more felinine than females (Hendriks et al. 2004). In cats but not dogs, cysteine is also used to generate felinine and isobuteine from isopentenyl pyrophosphate (formed from acetyl-CoA) and valine, respectively, via cooperation of the liver and kidneys (Che et al. 2021; Hendriks et al. 2008; RutherfurdMarkwick et al. 2005). Cysteine participates in all these reactions in the form of glutathione. These synthetic pathways may contribute to greater requirements of cats for cysteine, as compared with dogs (Li and Wu 2023a). Physiological functions of felinine, isovalthine, and isobuteine remain unknown but may serve as pheromones in cats for the purpose of territorial marking and intra-species communications (e.g., chemical signals to attract females) (Miyazaki et al. 2008).
Fig. 4.2 Synthesis of taurine from L-cysteine in the liver of dogs and cats. In the presence of serine, L-methionine is degraded on the liver via the transsulfuration pathway to generate cysteine and α-ketobutyrate. Subsequently, L-cysteine is oxidized to hypotaurine via reactions catalyzed by cysteine dioxygenase and cysteinesulfinate decarboxylase. Hypotaurine is spontaneously oxidized to taurine. Most breeds of dogs can synthesize sufficient taurine when fed L-methionine- or Lcysteine-adequate diets. However, this synthetic pathway is limited in cats due to low activities of cysteine dioxygenase and cysteinesulfinate decarboxylase
Fig. 4.2 Synthesis of taurine from L-cysteine in the liver of dogs and cats. In the presence of serine, L-methionine is degraded on the liver via the transsulfuration pathway to generate cysteine and α-ketobutyrate. Subsequently, L-cysteine is oxidized to hypotaurine via reactions catalyzed by cysteine dioxygenase and cysteinesulfinate decarboxylase. Hypotaurine is spontaneously oxidized to taurine. Most breeds of dogs can synthesize sufficient taurine when fed L-methionine- or Lcysteine-adequate diets. However, this synthetic pathway is limited in cats due to low activities of cysteine dioxygenase and cysteinesulfinate decarboxylase

4.5.4 AA Imbalances and Antagonisms, as Well as Metabolic Adaptation to AA Intakes

Compared with non-collagen animal proteins, plant proteins (particularly those in cereals) generally contain low amounts of nearly all AAs, particularly the AAs (e.g., lysine, tryptophan, threonine, methionine, and cysteine) that are not synthesized de novo and the AAs (glycine and proline + hydroxyproline) that are most abundant in the body (Li et al. 2021d). Dietary imbalances of AAs (defined as improper ratios of AAs in diets) can occur in dogs and cats fed plant-based diets containing a small amount of or no animal-sourced ingredients (Li and Wu 2023a). Antagonisms among AAs with similar chemical structures (e.g., BCAAs) or net electric charges (e.g., arginine and lysine) that share the same transmembrane transporters can occur in response to their improper ratios in diets or tissue (Wu 2022). Effects of dietary AA imbalances or antagonisms in dogs are more similar to those in rats than in cats based on alterations in food intake, plasma metabolic patterns, and hepatic enzyme activities (Morris 2002). Animal-sourced foods are abundant providers of all proteinogenic AAs as well as taurine and 4-hydroxyproline and play an important role in balancing AAs in diets for dogs and cats (Che et al. 2021; Li and Wu 2018; 2020; Li et al. 2021a; Wu et al. 2016a).
Dogs and cats have a high ability to use and oxidize dietary AAs, with the rate of AA catabolism in cats being greater than that in dogs under both fed and postabsorptive conditions (Morris 2002; Russell et al. 2002; Wester et al. 2015). For example, dogs and cats can tolerate at least 32% and 60% dietary protein, respectively (DeNapoli et al. 2000; Dodman et al. 1996; Green et al. 2008; Laflamme and Hannah 2013). These animals, just like most of other mammals (e.g., pigs and rats) can adapt to: (1) low-protein diets by increasing food intake and reducing whole-body AA catabolism, and (2) high-protein diets by reducing food intake and up-regulating the expression of AA-catabolic enzymes (Harper et al. 1984; Rogers et al. 1998). However, dogs are not as efficient as rats in metabolic adaptation to low- or highprotein intakes. In support of this view, studies with adult dogs have shown that the rate of whole-body protein degradation is not affected by low dietary protein intake and that the rate of whole-body leucine oxidation is not influenced by high dietary protein intake (e.g., increasing from 32 to 148 g CP/Mcal ME) (Humbert et al. 2002). In contrast, cats rapidly respond to protein-free diets by decreasing the rate of whole-body AA oxidation (Hendricks et al. 1996) and to elevated protein intake by increasing the rate of whole-body AA oxidation (Hendricks et al. 1997; Wester et al. 2015).

4.5.5 AA Requirements of Dogs and Cats

Compared with dogs, cats have greater a loss of endogenous nitrogen (urinary plus fecal nitrogen) when fed nitrogen-free diets (Hendriks et al. 1996 and 2002; Kendall et al. 1982). The whole-body degradation of protein provides AAs (e.g., arginine) for metabolic utilization. Accordingly, cats have greater requirements for dietary AAs than dogs (Mansilla et al. 2018; 2020a,b; NRC 2006). Quantitative requirements of dogs and cats for dietary AAs may be estimated from experiments involving nitrogen balance, the factorial analysis of use of AAs for growth and product (e.g., milk and hair) formation, and the oxidation of direct and indicator AAs (Lambie et al. 2024; Singh et al. 2024; Wu 2022). Interestingly, differences in dietary requirements for EAAs between these two animal species do not appear to be substantial (Rogers and Morris 1979). It is possible that cats degrade NEAAs at greater rates in a tissue-specific manner, as compared with dogs, with some of NEAAs being used for hepatic and renal gluconeogenesis. This is physiologically and nutritionally important because feline diets contain only a small amount of glycogen or starch (Washizu et al. 1999).
Skeletal muscle is the largest tissue in healthy animals. Like other mammals (e.g., humans), elderly dogs and cats experience sarcopenia (aging-related progressive losses of skeletal muscle mass and strength) because the rate of protein synthesis is lower than the rate of proteolysis (Kealy 1999; Laflamme 2008a,b). Sarcopenia results in not only physical weakness and disability but also tremendous increases in risks for falls, fractures, morbidity, and mortality, as well as health care costs. Exercise and AAs (e.g., arginine, branched-chain AAs, glutamine, glycine, and tryptophan) are co-stimuli to activate the mechanistic target of rapamycin cell signaling pathway and protein synthesis in skeletal muscle, thereby mitigating the loss of muscle protein with aging in humans (Glynn et al 2010; Holowaty et al. 2023; Mitchell et al. 2015). This is true for dogs and cats (Laflamme and Hannah 2013; Laflamme and Danièlle 2014; Williams et al. 2001), although the underlying signal transduction mechanisms have not yet been elucidated. There are suggestions that dietary protein requirements be increased by approximately 50% and 2.5-fold in older dogs and cats, respectively, compared with young adults (Laflamme 2008b; Laflamme and Hannah 2013; Wannemacher and McCoy 1966). This is equivalent to 2.55 and 5 g protein/kg BW/day for healthy adult dogs and cats, respectively (Churchill and Eirmann 2021). Considering long-term kidney health, we (Li and Wu 2023a) recommended adequate intakes of high-quality protein (i.e., 32% and 40% animal protein in diets for aging dogs and cats, respectively; DM basis) to alleviate sarcopenia. This goal can be achieved through the inclusion of pet-food grade animal-sourced foodstuffs in canine and feline diets (Li et al. 2021a, d; Li and Wu 2022, 2023b).

4.6 Vitamin Metabolism in Dogs and Cats

4.6.1 Vitamin Metabolism in Dogs

Like most animals, dogs can convert: (1) dietary β-carotene into vitamin A (alltrans-retinol) in enterocytes of the small intestine (Schweigert 1998), (2) tryptophan into niacin in the liver (Krehl 1981), and (3) glucose into vitamin C in the liver (McDowell 1989). Thus, β-carotene and tryptophan can replace some vitamin A and niacin in canine diets, respectively, and dogs do not have a dietary requirement for vitamin C. However, in contrast to other mammals (e.g., sheep, cattle, horses, pigs, rats, and humans), dogs have a limited ability to synthesize vitamin D3 (cholecalciferol) from 7-dehydrocholesterol in the skin exposed to sunlight and, therefore, depend on the dietary intake of vitamin D3 for bone health (How et al. 1994; Morris 1999). This is because the skin of dogs has a low concentration of 7-dehydrocholesterol and a high activity of 7-dehydrocholesterol ∆7 reductase to rapidly convert 7-dehydrocholesterol into cholesterol (Zafalon et al. 2020a). For example, concentrations of 7-dehydrocholesterol in the skin of dogs are only about 10% of those in the skin of rats (How et al. 1994). Cholecalciferol is generally added to the commercial diets for dogs, although dogs can effectively use the plantsourced vitamin D2 (ergocalciferol). Synthetic vitamin K3 (menadione sodium bisulfite; water-soluble) is used in the commercial diets of dogs. Dogs do not synthesize other vitamins required for cell metabolism (NRC 2006). Lipid- and water-soluble vitamins are excreted from the body mainly via bile (feces) and urine, respectively (Wu 2018).

4.6.2 Vitamin Metabolism in Cats

Vitamin B6, Vitamin B12, folate, and choline, along with methionine, serine, and histidine, play an important role in one-carbon metabolism in cats, as in other animals (Verbrugghe and Bakovic 2013). Like dogs, cats synthesize vitamin C and do not require a dietary source of this vitamin under normal feeding and environmental conditions (McDowell 1989), and also have a very limited ability to synthesize vitamin D3 from 7-dehydrocholesterol in response to sunlight (How et al. 1994). This is because of a very low concentration of 7-dehydrocholesterol and a high activity of 7-dehydrocholesterol ∆7 reductase to rapidly convert 7-dehydrocholesterol into cholesterol (How et al. 1994; Morris 1999), as reported for dogs. Cholecalciferol is generally added to the commercial diets for cats. Unlike dogs, cats cannot convert dietary β-carotene into vitamin A due to the absence of β-carotene 15,15-dioxygenase (Gershoff et al. 1957), although β-carotene is absorbed by the enterocytes of the small intestine (Schweigert et al. 2002). In addition, cats cannot use dietary vitamin D2 (the vitamin D in sun-dried plants). Furthermore, cats cannot synthesize niacin from tryptophan due to (a) a high activity of picolinate carboxylase to divert 2- amino-3-carboxymuconate semialdehyde into α-ketoadipate, and (b) a low activity of quinolinate phosphoribosyltransferase for converting 2-amino-3-carboxymuconate semialdehyde into nicotinate mononucleotide (Badawy 2017). Therefore, feline diets must provide vitamin A, vitamin D3, and niacin. Other vitamins are not synthesized by cats and are required in their diets for cell metabolism (NRC 2006).

4.6.3 Anti-vitamin Factors in Foods

Some foods contain anti-vitamin factors. For example, thiaminase is present in the viscera of some freshwater fish [e.g., common bream (Abramis brama), central stoneroller (Campostoma anomalum), goldfish (Carassius auratus), and common carp (Cyprinus carpio)], some marine fish [e.g., broad-striped anchovy (Anchoa hepsetus), Atlantic herring (Clupea harrengus), Atlantic cod (Gadus morhua), and Yellowfin tuna (Neothunnus macropterus)], and some bacteria (Greig and Gnaedinger 1971). This enzyme is heat-labile and is inactivated by cooking. Among the 32 species of freshwater fish tested, 18 of them were found to contain thiaminase. Among the 61 marine fish tested, 32 were found to contain thiaminase (Lichtenberger 2021). In other words, 56% of the freshwater fish examined contained thiaminase as compared with 51% of the marine fish species. Interestingly, there is little or a small difference in thiaminase activity between freshwater and marine fish. Furthermore, raw egg white contains avidin (a glycoprotein) that binds biotin very tightly. Fortunately, avidin is heat-labile and can be inactivated by cooking. Caution should be taken when feeding dogs and cats with raw fish or raw eggs.

4.7 Mineral Metabolism in Dogs and Cats

4.7.1 Mineral Metabolism in Dogs

In dogs, nutritionally essential macrominerals are Ca, P, Na, Cl, K, and Mg, whereas nutritionally essential microminerals are Fe, Zn, Cu, Mn, I, and Se (NRC 2006). Possibly essential minerals for these animals are Mo, boron, and Cr. Methionine and cysteine provide the sulfur atom required for the structures of protein and other molecules in dogs. Dietary inorganic sulfur is not required by these animals. Minerals are essential for nutrient transport, bone growth and development, the activities of proteins (including enzymes), and acid–base balance (Wu 2018). However, excessive Ca intake (e.g., > 0.31% of dietary DM) results in skeletal abnormalities in growing dogs (Hazewinkel et al. 1991; Hedhammar et al. 1974). Likewise, excessive Ca and P intake (0.31% Ca and 0.28% P of dietary DM) during early maturation (3 to 17 weeks of age) in dogs alters Ca and P balances by as early as 9 weeks of age, but dietary normalization (0.1% Ca and 0.08% P of dietary DM) during 17 to 27 weeks of age can effectively mitigate any long-term adverse effects on Ca and P balance (Schoenmakers et al. 1999).
Dogs ingesting 4 g NaCl/kg BW/day showed no sign of appreciable salt retention for 6 days if they have free access to drinking water (Ladd and Raisz 1949). However, seizures, hypernatremia [serum sodium concentration (211 mM; reference range, 145–158 mM) and serum chloride concentration (180 mM; reference range, 105– 122 mM), and death occur after consuming an excessive amount of sodium from a salt–flour mixture (finely ground salt; Khanna et al. 1997). Dogs consuming excessive salts should have free access to sufficient drinking water.

4.7.2 Mineral Metabolism in Cats

Cats have the same requirements for minerals as dogs but can tolerate a higher intake of dietary salt than dogs (NRC 2006). Cats are commonly beset with the problem of urolithiasis, namely the formation of sediment (consisting of one or more poorly soluble urine crystalloids) within the urinary tract (Houston et al. 2003). Magnesium ammonium phosphate (also known as struvite) and calcium oxalate (CaOx) are often present in the majority of such deposits in alkaline urine (Robertson et al. 2002). In North America, struvite and CaOx stones are the most and second most common mineral types found in the feline uroliths. Thus, caution must be exercised to avoid an excess of dietary magnesium and calcium. In practice, sufficient water intake can reduce risk for urolithiasis in cats. In addition, a high dietary intake of phosphorus and a low ratio of calcium-to-phosphorus in diets increases risks for kidney damage in cats (Summers et al. 2020). Furthermore, the content of minerals in the diet can affect its palatability for cats and must be monitored carefully. Of note, there is evidence that current commercial vegan foods for dogs and cats may not provide: (a) adequate calcium, potassium, and sodium for dogs; or (b) sufficient potassium and proper Ca/P ratio for cats (Zafalon et al. 2020b). This nutritional problem can be prevented by the inclusion of animal-sourced foodstuffs, which contain relatively a large amount of minerals (Li and Wu 2022, 2023b).

4.8 Water Metabolism in Dogs and Cats

4.8.1 Overview of Water Metabolism

Water is the quantitatively most important and most abundant nutrient for all animals, and its loss occurs via evaporation, respiration, urine, and feces (Wu 2018). Drinking and dietary (moist food) water are by far the major sources of water in animals, as the production of metabolic water from the oxidation of nutrients is limited in most mammals including dogs and cats (e.g., 1.09, 0.60, and 0.41 g water for the oxidation of 1 g fat, glucose, and protein, respectively (Wu 2018). Thirst results from both intracellular and extracellular dehydration (e.g., a loss of 0.5–1% of body water), as well as an increase in plasma osmolality, leading to the stimulation of water consumption (NRC 2006). It usually takes only days for animals (including dogs and cats) to be dehydrated particularly in an environment with elevated temperatures, but weeks or longer before clinical signs of a deficiency of a non-water nutrient are evident. In both dogs and cats, clinical signs of water deficiency include the loss of elasticity of the skin, reduced saliva secretion, reduced appetite, reduced blood volume, impaired blood circulation, the formation of bladder and kidney stones, and the dysfunction of tissues. These abnormalities appear when water stores in dogs are decreased by 5%, and death may occur when the body loses about 15% water (Kirk and Bistner 1981).

4.8.2 Water Metabolism in Dogs

Thirsty dogs can drink sufficient water within minutes to restore the losses of water, whereas starved dogs reduce water consumption by 67–80%. By contrast, increasing protein or salt intake increases water consumption in dogs. Voluntary water consumption by dogs is affected by the moisture of ingested foods. The ratio of total water intake to DM intake is 2.33 ml water for 1 g DM food. At 22–25 °C, a dog needs 50–60 ml water/kg BW/day (Schaer 1989). Water intake is increased when dogs exercise or are exposed to heat stress. Exposure to cold does not appear to affect water consumption.

4.8.3 Water Metabolism in Cats

Like dogs, cats require water as the most abundant nutrient and can maintain body water balance when fed meat containing ~ 70% water. At 22–25 °C, a cat needs 45–55 ml water/kg BW/day (NRC 2006). Cats have a lower physiological thirst drive than dogs and concentrate their urine (through the reabsorption of water in the lumen of kidney tubules into the blood) to a greater degree than dogs, which helps to conserve water (Anderson 1982; Chew 1965). In addition, the dense hair of cats can minimize the evaporative loss of water through the skin. Thus, cats may have a lower requirement for exogenous water than dogs under similar feeding conditions and can tolerate mild dehydration better than dogs (NRC 2006).
A loss of 20% water from cats can be fatal (Anderson 1982). Compared with continuous feeding, periodic feeding reduces the daily consumption of food and water by cats (Finco et al. 1986). In both dogs and cats, drinking too much water may result in adverse effects. As important as it is to avoid dehydration, water intoxication in dogs due to their excessive water intake may occur during swimming, diving, or water-retrieving (Flaim 2019). A cat drinking an excessive amount of water (e.g., > 100 ml/kg BW/day) may appear thirsty, but may actually have potential health problems (e.g., diabetes mellitus, chronic kidney disease, and hyperthyroidism; Sparkes et al. 2015).

4.9 Feeding Behavior and Management of Dogs and Cats

4.9.1 General Considerations

The evolution of domestic dogs and cats has resulted in the idiosyncrasies of their digestive tract, metabolism, and nutrient requirements. Both dogs and cats can develop learned taste aversions (Mugford 1977), avoid arginine-deficient diets (Morris 2002), prefer flavor components in meat and peptides (e.g., protein hydrolysates) and free AAs (e.g., alanine, histidine, leucine, lysine, and proline, as well as monosodium glutamate plus 5’-nucleotides (Hargrove et al. 1994; Kumazawa and Kurihara 1990a,b). In both animal species, age and the percentage of lean or fat body mass can influence food intake (Hall et al. 2018a, b). Their appetite is regulated by signals produced in the hypothalamus and peripheral organs (e.g., the digestive tract and white adipose tissue), as well as by the concentrations of glucose, fatty acids, amino acids, and their metabolites in the blood (Wu 2018). In addition, overweight or obesity occurs in dogs and cats if they are allowed to eat as much as they want at all times during the day without adequate exercise (Zoran 2010). Furthermore, spaying (the removal of ovaries from females) and neutering (the removal of testes from males) are performed as a management procedure before reaching sexual maturity (e.g., at or before 5 and 6 months of age in cats and dogs, respectively) with adequate nutritional support to prevent any unwanted behaviors (Hart et al. 2020; Vendramini et al. 2020). In both dogs and cats, care must be taken to prevent gastrointestinal parasites, as well as infectious diseases [e.g., canine and feline parvovirus (highly contagious viral diseases of dogs and cats, respectively) that cause acute gastrointestinal illness] (Laflamme and Hannah 2010). Of particular note, cow or goat milk is not a suitable replacement milk for both puppies and kittens, because the ruminant milk contains much less protein, arginine, taurine, methionine, fat, ME, calcium, phosphorus, sodium, copper, iron, and oleic acid than dog and cat milk (Debraekeleer et al. 2010; Gross et al. 2010), as shown in Table 4.3. Orphaned puppies and kittens can be fed commercial milk replacers especially produced for them. However, despite the above shared similarities, dogs and cats differentially select diets containing different compositions of macronutrient when given choices to eat foods with similar palatability (Hall et al. 2018a). These peculiar aspects of canine and feline nutrition are highlighted in the following paragraphs. Such knowledge will aid in our understanding of the different feeding behaviors of dogs and cats. This, in turn, can help to develop optimal management methods to raise, nurture, and care for them.
Table 4.3 Concentrations of nutrients in the milks of cats, dogs, cows, goats, and pigsa
Table 4.3 Concentrations of nutrients in the milks of cats, dogs, cows, goats, and pigsa

4.9.2 Weaning of Dogs and Cats

Weaning refers to the full transition of the puppy’s diet from its mother’s liquid milk to the solid growth diet. This is a significant and stressful period for young dogs. In dogs, weaning can begin at 3–4 weeks of age (Case et al. 2011). A small amount of canned pet food (usually consisting of ingredients of animal origin, such as meat, poultry, fish, and animal by-products) mixed with water (1:1, g/g) can be introduced to puppies 1 week before weaning so that they can adapt more easily to solid foods. Care must be taken to protect the intestinal microflora and prevent intestinal disorders in post-weaning dogs, such as gastritis, abnormal stool, gut atrophy, diarrhea, enteritis, leaky gut syndrome, inflammatory bowel disease, and vomiting (Case et al. 2011). The weanling diet must contain sufficient proteinogenic AAs (particularly arginine, glutamate, glutamine, glycine, methionine, proline, and tryptophan) and taurine (Wu 2018).
Weaning of cats usually begins when kittens are ~ 4 weeks of age by separating them from their mother for a few hours at a time each day and is completed at ~ 8 weeks of age (Case et al. 2011). Moistened kitten food and clean water bowls can be provided to kittens before the nursing phase ends, so as to stimulate the development of their digestive tract and facilitate their transition from the mother’s liquid milk to solid weaning diets. As for dogs, weaning allows kittens to gradually progress from the sole maternal care to social independence. This is a critical period in neonatal life when environmental stresses should be minimized or prevented. During weaning, cats must be provided with diets containing high-quality protein and taurine to ensure optimal nutrition for the small intestine and the whole body.

4.9.3 Food Selection of Dogs and Cats

Dogs are similar to other carnivores in the GIT anatomy but to omnivores in the metabolism of most nutrients. This characteristic is useful to guide canine feeding. Dogs have evolved in their ability to digest a relatively large amount of cooked starch (Axelsson et al. 2013; Félix et al. 2012) and select a diet lower in protein (30% of ME from protein) than a high-protein (52% of ME from protein) diet of wild wolves (Buff et al. 2014). Thus, dogs have more latitude in the selection of food ingredients and more flexible adaptability to both animal- and plant-sourced diets (NRC 2006). In addition, dogs have receptors for sweet substances and, therefore, have preference for 10% sucrose solution, compared to water. When offered complete and balanced diets with varying levels of protein, fat, and carbohydrate, dogs will choose a high-fat (63% of ME from fat) diet over a low-fat (15% of ME from fat) diet (Bradshaw 2006; Buff et al. 2014; Hewson-Hughes et al., 2016). Regarding dietary protein choices, dogs are more like rats than cats. When offered various levels of protein in diets, dogs will select the diets containing 25–30% of energy as protein (Romos and Ferguson 1983; Tôrres et al. 2003). Dogs can eat powdered, dry, semi-moist, or canned diets. Food should be stored in under air-tight, dry, and cool conditions.
Cats are picky eaters (Pekel et al. 2020). Preferring meat with palatability factors (e.g., fats, AAs, peptides, and nucleotides) as food, cats eat small prey, such as rats, mice, birds, lizards, and insects. Odor, taste, texture, particle size, and temperature of diets affect the food preference of cats. Cats may avoid a particular diet because of the textural difference but not the difference in its nutrient composition. Clearly, cats prefer (1) moist foods to dry foods, (2) warm foods to cold or hot foods, (3) highprotein (e.g., 50% protein on a DM basis) foods to low-protein (e.g., 20% protein on a DM basis) or protein-free foods, (4) animal fat to plant fat, and (5) diets containing 25% total lipids to diets containing 10% or 50% total lipids (Cook et al. 1985, 1996; Eyre et al. 2022; Morris 2002; Rogers et al. 2004). Thus, cats would select beef tallow over butter and chicken fats, but have similar preference for beef tallow, lard, and partially hydrogenated vegetable oil (Kane et al. 1981). A diet consisting of high fat content and 5% hydrogenated beef tallow is particularly palatable to cats (Rogers et al. 2004). By contrast, medium-chain fatty acids, present in either the free form (e.g., 0.1% caprylic acid) or TAG (e.g., 5% medium-chain TAG), and 25% hydrogenated coconut oil in diets can reduce food intake by cats (MacDonald et al. 1985). Of particular note, cats cannot taste sweetness as indicated previously, because of the lack of sweet taste receptors due to the deletion of the Tas1r gene and, unlike dogs, do not select sweet substances such as sucrose (Li et al. 2005, 2006). For this reason, cats show no preference for sugar-rich foods such as fruits and juice.
When cats are offered complete and balanced diets with varying levels of protein, fat, and carbohydrate (without controlling palatability), cats will choose high-protein diets (Bradshaw 2006; Hewson-Hughes et al. 2016). Thus, high-starch diets over a prolonged period of time could be a major contributing to poor feline health and must be avoided in feeding both young and adult cats particularly those with overweight or diabetes (Verbrugghe and Hesta 2017). Likewise, domestic cats select a macronutrient profile (52% and 48% of ME from protein and fat, respectively) similar to the diet (52%, 46%, and 2% of ME from protein, fat, and glycogen, respectively) of wild cats. With regards to AAs, cats have preference for proline, cysteine, ornithine, lysine, histidine, alanine, and leucine (Bradshaw et al. 1996), but against arginine, isoleucine, phenylalanine, and tryptophan (Oliveira et al. 2016; Zaghini and Biagi 2005). This is consistent with the notion that cats are more sensitive to bitter taste than dogs. Exposure of kittens to numerous different flavors and textures early in life can result in wider preference for such foods later in life (Kuo 1967). However, adult cats can learn to select new and better flavors in addition to the ones early in life (Mugford 1977).
Healthy cats do not select sodium even if they are deficient in sodium, but acidotic cats will select diets containing excess sodium (Yu et al. 1997). Furthermore, cats can develop a learned taste aversion to diets that causes metabolic acidosis (Cook et al. 1996). Healthy cats are normally neophilic, namely selecting a new food (or flavor) if offered a variety of foods (or flavors) throughout their lives (Beauchamp et al. 1977). Furthermore, cats do not consume a powdered diet but will ingest the same diet in a pelleted or gel form. If a diet is too dry or powdery, cats will eat very little even when no alternative food is offered (NRC 2006).

4.9.4 Meal Frequency of Dogs and Cats

Dogs eat small and large prey, as well as plant-sourced foods. These animals can consume a large meal within 10 min that is sufficient to meet daily nutrient requirements, but they usually eat 4 to 8 meals/day (depending on breed; Bradshaw 2006). Growing puppies should have free access to food or be fed 2–3 times daily. Young and adult dogs generally ingest food and drink water during the light period and may eat at night if they are hungry or dehydrated, but some breeds of dogs also eat and drink during the dark period (Mugford and Thorne 1980). On average, dogs drink more water than cats per kg BW per day. Although adult dogs may do well on a regimen of eating a meal per day with possible health benefits (Bray et al. 2021), these animals should be offered foods at least twice daily to ensure adequate nutrition and welfare (Brooks et al. 2014). Multiple meals per day may help dogs to alleviate boredom, while reducing the risk for the problem of gastric dilatation-volvulus in susceptible breeds of dogs (Bataller 1995). If overweight or obese, they should be fed either less or lower energy-dense diets. Caloric restriction may reduce oxidative stress and delay aging in dogs (Kealy et al 2002; Lawler et al. 2008). Of particular note, intermittent fasting on a ketogenic diet containing medium-chain triglycerides may confer both metabolic and immunological effects in healthy dogs (Leung et al. 2020). Dogs with special health issues or dietary needs may require individualized feeding programs.
In contrast to dogs, cats do not have clear-cut circadian rhythms (Hawking et al. 1971; Randall et al. 1985). Some cats are nocturnal (e.g., resting during the daytime and going out or walking at night), and their existing rhythms can be altered by minor disturbances such as human noises (Macdonald and Apps 1978). Cats generally eat frequent, small meals and drink water throughout the day and night and can voluntarily eat 12–20 meals/day, but fresh food should be offered daily (Kane et al. 1981). After they adapt to a new diet for 1 week, the size and number of meals are not affected by the type of diet. When cats eat a big portion of food at one time, they may immediately vomit it due to gastric irritation. Therefore, appropriate and frequent meals are critical for feline health. Pet owners can decide meal size and frequency based on breed, age, health condition, ambient environment, and practicality for their dogs and cats rather than simply food availability.
Camara et al. (2020) recently reported that compared with the feeding of the same canned food four times/day, adult cats fed only once daily had greater plasma concentrations of glucagon-like protein-1, gastric inhibitory protein, and some AAs within the initial postprandial hours (as expected because of more nutrient intakes at a single time), but a lower value of respiratory quotient (RQ) in the fasting state. These authors interpreted the data as that a reduced RQ value may result from a greater rate of whole-body fatty acid oxidation and that feeding once daily may be a beneficial feeding management strategy for indoor cats to promote satiation and lean body mass. However, a reduced RQ value in the fasting state may also result from a greater rate of the conversion of AAs (released from the breakdown of muscle protein; possibly an adverse effect) into glucose, because this process has an RQ value of 0.333 (Wu 2018). In the work (Study 1) of Camara et al. (2020), neither body weight nor feed (energy) intake differed between the two groups of cats. Clearly, more studies are warranted to assess whether feeding once daily may have health benefits for cats.

4.10 Conclusion and Perspectives

Both dogs and cats naturally prefer and effectively digest meat, but can also digest appropriate amounts of properly (e.g., suitable temperature and period) processed and formulated plant-sourced ingredients. These two animal species differ markedly in some aspects of nutrition and metabolism: including (1) the ability to digest the quantities (i.e., small vs large) of dietary cooked starch in the small intestine; (2) thermoneutral zones and rates of basal energy metabolism; (3) qualitative (i.e., presence or absence) and quantitative (i.e., amounts) requirements for many dietary nutrients, particularly protein, certain AAs (arginine, taurine, methionine, and cysteine, as well as NEAAs), ω3 and ω6 PUFAs, and vitamins (e.g., vitamin A and niacin); (4) sensitivity to dietary AA imbalances and antagonisms; (5) felinine synthesis from cysteine; (6) feeding behavior and meal frequencies; (7) food preference and refusal; (8) the selection for foods containing various amounts of protein and fats; (9) the form of diets; and (10) the occurrence of gastrointestinal and metabolic disorders (Table 4.1). Recommended minimum requirements and allowances of dietary nutrients other than AAs for growing and adult dogs and cats are summarized in Table 4.4. Recommended values for their dietary AA requirements have been recently summarized (Li and Wu 2023a, b). Unlike cats, dogs have adapted by expressing a high activity of pancreatic α-amylase activity during evolution to omnivorous diets containing taurine-rich meat and plant ingredients with relatively a large amount of starch. Thus, a high intake of starch (e.g., 40% on a DM basis) can result in metabolic disorders in cats but not in dogs, whereas dietary fiber is beneficial for the intestinal health of both cats and dogs. Given the increasing incidences of obesity or overweight in both dogs and cats, dietary supplementation with arginine may be a promising solution to prevent metabolic syndromes in these animals, as reported for rats (Jobgen et al. 2009, 2022a, b, 2023). In practice, cats should not be fed canine foods. Because the composition of the milk of both dogs and cats differ from that of farm mammals, the young pets should not be fed replacer diets formulated based on goat or cow milk. For both dogs and cats, there may be breed differences in dietary requirements for nutrients and thermoneutral zones even among different sizes of the same breed. The fundamental knowledge of the idiosyncrasies of metabolism in dogs and cats is essential for guiding their feeding and care, as well as food manufacturing. Of particular note, current commercial vegan petfoods may be nutritionally inadequate for dogs (low content of calcium, potassium, sodium, and methionine) and cats (low content of protein, arginine, taurine, and potassium, as well as an improper Ca/P ratio). Animal-sourced foods contain nutritionally significant amounts of AAs, lipids, and minerals and therefore play an important role in balancing AAs in diets for dogs and cats play an important role in optimizing the nutrition and health of these companion animals.
Table 4.4 Recommended allowances of dietary fatty acids, minerals, and vitamins for dogs, cats, and pigs
Table 4.4 Recommended allowances of dietary fatty acids, minerals, and vitamins for dogs, cats, and pigs
      
This chapter was originally published in Nutrition and Metabolism of Dogs and Cats, Advances in Experimental Medicine and Biology 1446, https://doi.org/10.1007/978-3-031-54192-6_4. This is an Open Access chapter licensed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/).

Anderson HD, Johnson BC, Arnold A (1940) The composition of dog’s milk. Am J Physiol 129:631– 634

Anderson RS (1982) Water balance in the dog and cat. J Small Anim Pract 23:588–598

Anderson PJ, Rogers QR, Morris JG (2002) Cats require more dietary phenylalanine or tyrosine for melanin deposition in hair than for maximal growth. J Nutr 132:2037–2042

Axelsson E, Ratnakumar A, Arendt M-L et al (2013) The genomic signature of dog domestication reveals adaptation to a starch-rich diet. Nature 495:360–364

Backus RC, Cohen G, Pion PD, Good KL, Rogers QR, Fascetti AJ (2003) Taurine deficiency in Newfoundlands fed commercially available complete and balanced diets. J Am Vet Med Assoc 223:1130–1136

Backus RC, Ko KS, Fascetti AJ, Kittleson MD, MacDonald KA, Maggs DJ, Berg JR, Rogers QR (2006) Low plasma taurine concentration in Newfoundland dogs is associated with low plasma methionine and cyst(e)ine concentrations and low taurine synthesis. J Nutr 136:2525–2533

Badawy AA (2017) Kynurenine pathway of tryptophan metabolism: Regulatory and functional aspects. Int J Tryptophan Res 10:1178646917691938

Baker DH, Czarnecki-Maulden GL (1991) Comparative nutrition of cats and dogs. Annu Rev Nutr 11:239–263

Bartges JM, Cook M (1999) Influence of feeding conjugated linoleic acid on body composition in healthy adult cats. In: Proceedings 17th ACVIM. pp 729

Bataller N (1995) Risk factors and the debate of diet in canine gastric dilatation-volvulus. Vet Clin Nutr 2:87

Batchelor DJ, Al-Rammahi M, Moran AW, Brand JG, Li X, Haskins M, German AJ, ShiraziBeechey SP (2011) Sodium/glucose cotransporter-1, sweet receptor, and disaccharidase expression in the intestine of the domestic dog and cat: two species of different dietary habit. Am J Physiol 300:R67-75

Bauer JE (1997) Fatty acid metabolism in domestic cats (Felis catus) and cheetahs (Acinonyx jubatas). Proc Nutr Soc 56:1013–1024

Bauer JE (2007) Responses of dogs to dietary omega-3 fatty acids. J Am Vet Med Assoc 231:1657– 1661

Beauchamp GK, Maller O, Rogers JG Jr (1977) Flavor preferences in cats (Felis catus and Panthera sp). J Comp Physiol Psychol 91:1118–1127

Beliveau GP, Freedland RA (1982) Metabolism of serine, glycine and threonine in isolated cat hepatocytes Felis domestica. Comp Biochem Physiol B 71:13–18

Belo PS, Romsos DR, Leveille GA (1976) Influence of diet on glucose tolerance, on the rate of glucose utilization and on gluconeogenic enzyme activities in the dog. J Nutr 106:1465–1474

Bermingham EN, Thomas DG, Morris PJ, Hawthorne AJ (2010) Energy requirements of adult cats. Br J Nutr 103:1083–1093

Besch EL, Woods JE (1977) Heat production biorhythms of laboratory animals. Lab Anim Sci 27:54–59

Biourge V, Sergheraert R (2002) Hair pigmentation can be affected by diet in dogs. Proc Comp Nutr Soc 4:103–104

Bordoni A, Biagi PL, Turchetto E, Hrelia S (1988) Aging influence on delta-6-desaturase activity and fatty acid composition of rat liver microsomes. Biochem Int 17:1001–1009

Bradshaw JWS (2006) The evolutionary basis for the feeding behavior of domestic dogs (Canis familiaris) and cats (Felis catus). J Nutr 136:1927S-1931S

Bradshaw JWS, Goodwin D, Legrand-Defretin V, Nott HMR (1996) Food selection by the domestic cat, an obligate carnivore. Comp Biochem Physiol A 114:205–209

Bray EE, Zheng Z, Tolbert MK, McCoy BM, Dog Aging Project Consortium, Kaeberlein M, Kerr KF (2021) Once-daily feeding is associated with better health in companion dogs: results from the Dog Aging Project. bioRxiv preprint. https://doi.org/10.1101/2021.11.08.467616

Brody S, Procter RC, Ural S Ashworth US (1934) Growth and development with special reference to domestic animals. XXXIV, Basal metabolism, endogenous nitrogen, creatinine and neutral sulphur excretions as functions of body weight. Agric Exp Stn Res Bull No 220. University of Missouri, Columbia

Brooks D, Churchill J, Fein K et al (2014) AAHA weight management guidelines for dogs and cats. J Am Anim Hospital Assoc 50:1–11

Brosey BP, Hill RC, Scott KC (2000) Gastrointestinal volatile fatty acid concentrations and pH in cats. Am J Vet Res 61:359–361

Bruhn JC (2017) Dairy goat milk composition. https://drinc.ucdavis.edu/goat-dairy-foods/dairygoat-milk-composition. Accessed 5 Oct 2021

Bryan DL, Hart P, Forsyth K, Gibson R (2001) Incorporation of alpha-linolenic acid and linoleic acid into human respiratory epithelial cell lines. Lipids 36:713–717

Buff PR, Carter RA, Bauer JE, Kersey JH (2014) Natural pet food: a review of natural diets and their impact on canine and feline physiology. J Anim Sci 92:3781–3791

Burns RA, Milner JA, Corbin JE (1981) Arginine: An indispensable amino acid for mature dogs. J Nutr 111:1020–1024

Camara A, Verbrugghe A, Cargo-Froom C, Hogan K, DeVries TJ, Sanchez A, Robinson LE, Shoveller AK (2020) The daytime feeding frequency affects appetite-regulating hormones, amino acids, physical activity, and respiratory quotient, but not energy expenditure, in adult cats fed regimens for 21 days. PLoS ONE 15:e0238522

Carciofi AC, Takakura FS, de-Oliveira LD, Teshima E, Jeremias JT, Brunetto MA, Prada F (2008) Effects of six carbohydrate sources on dog diet digestibility and post-prandial glucose and insulin response. J Anim Physiol Anim Nutr 92:326–336

Case LP, Daristotle L, Hayek MG, Raasch MF (2011) Canine and feline nutrition, 3rd edn. Mosby, Maryland Heights, Missouri

Cavanaugh SM, Cavanaugh RP, Gilbert GE, Leavitt EL, Ketz is JK, Vieira AB (2021) Shortterm amino acid, clinicopathologic, and echocardiographic findings in healthy dogs fed a commercial plant-based diet. PLoS ONE 16:e0258044

Che DS, Nyingwa PS, Ralinala KM, Maswanganye GMT, Wu G (2021) Amino acids in the nutrition, metabolism, and health of domestic cats. Adv Exp Med Biol 1285:217–231

Chew RM (1965) Water metabolism of mammals. In: Mayer WV, van Gelder RG (eds) Physiological mammalogy: mammalian reactions to stressful environments. Academic Press, New York, pp 43–178

Chiocchetti R, Galiazzo G, Giancola F, Tagliavia C, Bernardini C, Forni M, Pietra M (2021) Localization of the serotonin transporter in the dog intestine and comparison to the rat and human intestines. Front Vet Sci 8:802479

Churchill JA, Eirmann L (2021) Senior pet nutrition and management. Vet Clin North Am Sm Anim Pract 51:635–651

Clark MH, Ferguson DC, Bunick D, Hoenig M (2013) Molecular and histological evidence of brown adipose tissue in adult cats. Vet J 195:66–72

Cook NE, Kane E, Rogers QR, Morris JG (1985) Self selection of dietary casein and soy-protein by the cat. Physiol Behav 34:583–594

Cook NE, Rogers QR, Morris JG (1996) Acid-base balance affects dietary choice in cats. Appetite 26:175–192

Croniger CM, Olswang Y, Reshef L, Kalhan SC, Tilghman SM, Hanson RW (2002) Phosphoenolpyruvate carboxykinase revisited: insight into its metabolic role. Biochem Mol Bil Educ 30:14–20

Czarnecki GL, Baker DH (1984) Urea cycle function in the dog with emphasis on the role of arginine. J Nutr 114:581–590

Davis TA, Nguyen HV, Garcia-Bravo R, Fiorotto ML, Jackson EM, Lewis DS, Lee DR, Reeds PJ (1994a) Amino acid composition of human milk is not unique. J Nutr 124:1126–1132

Davis TA, Nguyen HV, Garcia-Bravo R, Fiorotto ML, Jackson EM, Reeds PJ (1994b) Amino acid composition of the milk of some mammalian species changes with stage of lactation. Br J Nutr 72:845–853

De Bruijne JJ, van den Brom WE (1986) The effect of long-term fasting on ketone body metabolism in the dog. Comp Biochem Physiol B 83:391–395

Debraekeleer J, Gross KL, Zicker SC (2010) Feeding nursing and orphaned puppies from birth to weaning. In: Thatcher CD, Remillard RL, Roudebush P (eds) Small Animal clinical nutrition Hand MS. Mark Morris Institute, Topeka, KS, pp 295–309

DeNapoli JS, Dodman NH, Shuster L, Rand WM, Gross KL (2000) Effect of dietary protein content and tryptophan supplementation on dominance aggression, territorial aggression, and hyperactivity in dogs. J Am Vet Med Assoc 217:504–508

Dillon EL, Wu G (2021) Cortisol enhances ctrulline synthesis from proline in enterocytes of suckling piglets. Amino Acids 53:1957–1966

Dobenecker B, Zottmann B, Kienzle E, Zentek J (1998) Investigations on milk composition and milk yield in queens. J Nutr 128:2618S-2619S

Dodman NH, Reisner I, Shuster L, Rand W, Luescher UA, Robinson I, Houpt KA (1996) Effect of dietary protein content on behavior in dogs. J Am Vet Med Assoc 208:376–379

Dunbar BL, Bauer JE (2002) Conversion of essential fatty acids by delta-6 desaturase in dog liver microsomes. J Nutr 132:1701S-1703S

Eyjolfson V, Spriet LL, Dyck DJ (2004) Conjugated linoleic acid improves insulin sensitivity in young, sedentary humans. Med Sci Sports Exerc 36:814–820

Eyre R, Trehiou M, Marshall E, Carvell-Miller L, Goyon A, McGrane S (2022) Aging cats prefer warm food. J Vet Behav 47:86–92

Félix AP, Gabeloni LR, Brito CBM, Oliveira SG, Silva AVF, Maiorka A (2012) Effect of βmannanase on the digestibility of diets with different protein sources in dogs determined by different methodologies. J Anim Sci 90:3060–3067

Feng BC, Li J, Kliegman RM (1996) Transcription of hepatic cytosolic phosphoenolpyruvate carboxykinase gene in newborn dogs. Biochem Mol Med 59:13–19

Finco DR, Adams DD, Crowelle WA, Stattelman AJ, Brown SA, Barsanti JA (1986) Food and water intake and urine composition in cats: influence of continuous versus periodic feeding. Am J Vet Res 47:1638–1642

Flaim D (2019) Can dogs drink too much water? The dangers of water intoxication. American Kennel Club. https://www.akc.org/expert-advice/health/can-dogs-drink-much-water-dangers-water-int oxication. Accessed 12 Jan 2022

Galim EB, Hruska K, Bier DM, Matthews DE, Haymond MW (1980) Branched-chain amino acid nitrogen transfer to alanine in vivo in dogs. J Clin Invest 66:1295–1304

Gershoff SN, Andrus SB, Hegsted DM, Lentini EA (1957) Vitamin A deficiency in cats. Lab Invest 6:227–240

Gil F, Arencibia A, García V, Ramírez G, Vázquez JM (2018) Anatomic and magnetic resonance imaging features of the salivary glands in the dog. Anat Histol Embryol 47:551–559

Glynn EL, Fry CS, Drummond MJ, Timmerman KL, Dhanani S, Volpi E, Rasmussen BB (2010) Excess leucine intake enhances muscle anabolic signaling but not net protein anabolism in young men and women. J Nutr 140:1970–1976

Green AS, Ramsey JJ, Villaverde C, Asami DK, Wei A, Fascetti AJ (2008) Cats are able to adapt protein oxidation to protein intake provided their requirement for dietary protein is met. J Nutr 138:1053–1060

Greig RA, Gnaedinger RH (1971) Occurrence of thiaminase in some common aquatic animals of the United States and Canada. Special Scientific Report—Fish. US Dept Commer Natl Mar Fish Serv 631:1–7

Gross KL, Becvarova I, Debraekeleer J (2010) Feeding nursing and orphaned kittens from birth to weaning. In: Thatcher CD, Remillard RL, Roudebush P (eds) Small Animal clinical nutrition Hand MS. Mark Morris Institute, Topeka, KS, pp 415–427

Hall JA, Vondran JC, Vanchina MA, Jewell DE (2018a) When fed foods with similar palatability, healthy adult dogs and cats choose different macronutrient compositions. J Exp Biol 221:jeb173450

Hall JA, Jackson MI, Vondran JC, Vanchina MA, Jewell DE (2018b) Comparison of circulating metabolite concentrations in dogs and cats when allowed to freely choose macronutrient intake. Biol Open 7:bio036228

Hammer VA, Rogers QR, Freedland RA (1996) Threonine is catabolized by L-threonine 3- dehydrogenase and threonine dehydratase in hepatocytes from domestic cats (Felis domestica). J Nutr 126:2218–2226

Hanson RW, Garber AJ (1972) Phosphoenolpyruvate carboxykinase. I. Its role in gluconeogenesis. Am J Clin Nutr 25:1010–1021

Hargrove DM, Morris JG, Rogers QR (1994) Kittens choose a high leucine diet even when isoleucine and valine are the limiting amino acids. J Nutr 124:689–693

Harper AE, Miller RM, Block KP (1984) Branched-chain amino acid metabolism. Annu Rev Nutr 4:409–454

Hart BL, Hart LA, Thigpen AP, Willits NH (2020) Assisting decision-making on age of neutering for 35 breeds of dogs: associated joint disorders, cancers, and urinary incontinence. Front Vet Sci 7:388

Hawking F, Lobban MC, Gammage K, Worms MJ (1971) Circadian rhythms (activity, temperature, urine and microfilariae) in dog, cat, hen, duck, thamnomys and gerbillus. Int J Cycle Res 2:455– 473

Hazewinkel HAW, van den Brom WE, van’t Klooster A, Voorhout G, van Wees A (1991) Calcium metabolism in Great Dane dogs fed diets with various calcium and phosphorus contents. J Nutr 121:S99–S106

He W, Connolly ED, Wu G (2024) Characteristics of the digestive tract of dogs and cats. Adv Exp Med Biol 1446:15–38. https://doi.org/10.1007/978-3-031-54192-6_2

He WL, Wu G (2022) Oxidation of amino acids, glucose, and fatty acids as metabolic fuels in enterocytes of developing pigs. Amino Acids 54:1025–1039

He WL, Posey EA, Steele CC, Savell JW, Bazer FW, Wu G (2023) Dietary glycine supplementation enhances postweaning growth and meat quality of pigs with intrauterine growth restriction. J Anim Sci 101:skadskad354

Hedhammar A, Wu FM, Krook L, Schryver HF, de LaHunta A, Wahlen JP, Kallfelz FA, Nunez EA, Hintz HF, Sheffy BE, Ryan GD (1974) Overnutrition and skeletal disease. An experimental study in growing Great Dane dogs. Cornell Vet 64 (suppl 5):1–160

Heinze CR, Larsen JA, Kass PH, Fascetti AJ (2009) Plasma amino acid and whole blood taurine concentrations in cats eating commercially prepared diets. Am J Vet Res 70:1374–1382

Heinze CR, Freeman LM, Martin CR, Power ML, Fascetti AJ (2014) Comparison of the nutrient composition of commercial dog milk replacers with that of dog milk. J Am Vet Med Assoc 244:1413–1422

Hendriks WH, Moughan PJ, Tarttelin MF (1996) Gut endogenous nitrogen and amino acid excretions in adult domestic cats fed a protein-free or an enzymatically hydrolyzed casein-based diet. J Nutr 126:955–962

Hendricks WH, Moughan PJ, Tarttelin MF (1997) Urinary excretion of endogenous nitrogen metabolites in adult domestic cats using a protein-free diet and the regression technique. J Nutr 127:623–629

Hendriks WH, Sritharan K, Hodgkinson SM (2002) Comparison of the endogenous ileal and faecal amino acid excretion in the dog (Canis familiaris) and the rat (Rattus rattus) determined under protein-free feeding and peptide alimentation. J Anim Physiol Anim Nutr 86:333–341

Hendriks WH, Vather R, Rutherfurd SM, Weidgraaf K, Rutherfurd-Markwick KJ (2004) Urinary isovalthine excretion in adult cats is not gender dependent or increased by oral leucine supplementation. J Nutr 134:2114S-2116S

Hendriks WH, Rutherfurd-Markwick KJ, Weidgraaf K, Morton RH, Rogers QR (2008) Urinary felinine excretion in intact male cats is increased by dietary cystine. Br J Nutr 100:801–809

Hewson-Hughes AK, Gilham MS, Upton S, Colyer A, Butterwick R, Miller AT (2011) The effect of dietary starch level on postprandial glucose and insulin concentrations in cats and dogs. Br J Nutr 106:S105–S109

Hewson-Hughes AK, Colyer A, Simpson SJ, Raubenheimer D (2016) Balancing macronutrient intake in a mammalian carnivore: disentangling the influences of flavour and nutrition. R Soc Open Sci 3:160081

Hill RC (1998) The nutritional requirements of exercising dogs. J Nutr 128:2686S-2690S

Hill RC, Scott KC (2004) Energy requirements and body surface area of cats and dogs. JAVMA 225:689–694

Holloway B, Stribling D, Freeman S, Jamieson L (1985) The thermogenic role of adipose tissue in the dog. Int J Obesity 9:423–432

Holowaty MNH, Lees MJ, Abou Sawan S, Paulussen KJM, Jäger R, Purpura M, Paluska SA, Burd NA, Hodson N, Moore DR (2023) Leucine ingestion promotes mTOR translocation to the periphery and enhances total and peripheral RPS6 phosphorylation in human skeletal muscle. Amino Acids 55:253–261

Hou YQ, Wu G (2017) Nutritionally nonessential amino acids: a misnomer in nutritional sciences. Adv Nutr 8:137–139

Hou YQ, Yin YL, Wu G (2015) Dietary essentiality of “nutritionally nonessential amino acids” for animals and humans. Exp Biol Med 240:997–1007

Hou YQ, Yao K, Yin YL, Wu G (2016) Endogenous synthesis of amino acids limits growth, lactation and reproduction of animals. Adv Nutr 7:331–342

Hou YQ, Wu ZL, Dai ZL, Wang GH, Wu G (2017) Protein hydrolysates in animal nutrition: Industrial production, bioactive peptides, and functional significance. J Anim Sci Biotechnol 8:24

Hou YQ, He WL, Hu SD, Wu G (2019) Composition of polyamines and amino acids in plant-source foods for human consumption. Amino Acids 51:1153–1165

Houston DM, Moore AEP, Favrin MG, Brent Hoff B (2003) Feline urethral plugs and bladder uroliths: a review of 5484 submissions 1998–2003. Can Vet J 44:974–977

How KL, Hazewinkel HAW, Mol JA (1994) Dietary vitamin D dependence of cat and dog due to inadequate cutaneous synthesis of vitamin D. Gen Comp Endocrinol 96:12–18

Hu SD, He WL, Wu G (2021) Hydroxyproline in animal metabolism, nutrition, and cell signaling. Amino Acids 54:513–528

Humbert B, Martin L, Dumon H, Darmaun D, Nguyen P (2002) Dietary protein level affects protein metabolism during the postabsorptive state in dogs. J Nutr 132:1676S – 1678

Hurley WL (2015) Composition of sow colostrum and milk. In: Farmer C (ed) The gestating and lactating sow. Wageningen Academic Publishers, Wageningen, The, Netherlands, pp 193–229

Iyer MS, Paszkiewicz RL, Bergman RN, Richey JM, Woolcott OO, Asare-Bediako I, Wu Q, Kim SP, Stefanovski D, Kolka CM, Clegg DJ, Kabir M (2019) Activation of NPRs and UCP1- independent pathway following CB1R antagonist treatment is associated with adipose tissue beiging in fat-fed male dogs. Am J Physiol 317:E535-547

Jacobsen JG, Smith LH (1968) Biochemistry and physiology of taurine and taurine derivatives. Physiol Rev 48:424–511

Jobgen WJ, Meininger CJ, Jobgen SC, Li P, Lee M-J, Smith SB, Spencer TE, Fried SK, Wu G (2009) Dietary L-arginine supplementation reduces white-fat gain and enhances skeletal muscle and brown fat masses in diet-induced obese rats. J Nutr 139:230–237

Jobgen WS, Wu G (2022a) L-Arginine increases AMPK phosphorylation and the oxidation of energy substrates in hepatocytes, skeletal muscle cells, and adipocytes. Amino Acids 54:1553–1568

Jobgen WS, Wu G (2022b) Dietary L-arginine supplementation increases the hepatic expression of AMP-activated protein kinase in rats. Amino Acids 54:1569–1584

Jobgen WS, Lee M-J, Fried SK, Wu G (2023) L-Arginine supplementation regulates energysubstrate metabolism in skeletal muscle and adipose tissue of diet-induced obese rats. Exp Biol Med 248:209–216

Kane E, Rogers QR, Morris JG, Leung PMB (1981) Feeding behavior of the cat fed laboratory and commercial diets. Nutr Res 1:499–507

Kaplan JL, Stern JA, Fascetti AJ, Larsen JA, Skolnik H, Peddle GD, Kienle RD, Waxman A, Cocchiaro M, Gunther-Harrington CT, Klose T, LaFauci K, Lefbom B, Lamy MM, Malakoff R, Nishimura S, Oldach M, Rosenthal S, Stauthammer C, O’Sullivan L, Visser LC, Williams R, Ontiveros E (2018) Taurine deficiency and dilated cardiomyopathy in golden retrievers fed commercial diets. PLoS ONE 13:e0209112

Kealy RD (1999) Factors influencing lean body mass in aging dogs. Comp Cont Edu Small Anim Pract 2:34–37

Kealy RD, Lawler DF, Ballam JM et al (2002) Effects of diet restriction on life span and 358 age-related changes in dogs. J Am Vet Med Assoc 220:1315–1320

Kendall PT, Blaze SE, Holme DW (1982) Assessment of endogenous nitrogen output in adult dogs of contrasting size using a protein-free diet. J Nutr 112:1281–1286

Kettelhut IC, Foss MC, Migliorini RH (1980) Glucose homeostasis in a carnivorous animal (cat) and in rats fed a high-protein diet. Am J Physiol 239:R437-444

Khanna C, Boermans HJ, Wilcock B (1997) Fatal hypernatremia in a dog from salt ingestion. J Am Anim Hosp Assoc 33:113–116

Kienzle E (1993a) Carbohydrate metabolism of the cat. 1. Activity of amylase in the gastrointestinal tract of the cat. J Anim Physiol Anim Nutr 69:92–101

Kienzle E (1993b) Carbohydrate metabolism of the cat. 2. Digestion of starch. J Anim Physiol Anim Nutr 69:102–114

Kirk RW, Bistner SI (1981) Handbook of veterinary procedures and emergency treatment. W.B. Saunders, Philadelphia, p 582

Kittleson MD, Keene B, Pion PD, Loyer CG (1997) Results of the multicenter spaniel trial (MUST): taurine- and carnitine-responsive dilated cardiomyopathy in American cocker spaniels with decreased plasma taurine concentration. J Vet Intern Med 11:204–211

Kleiber M (1961) The fire of life. Wiley & Sons, New York Knopf K, Sturman JA, Armstrong M, Hayes KC (1978) Taurine: an essential nutrient for the cat. J Nutr 108:773–778

Kompan D, Komprej A (2012) The effect of fatty acids in goat milk on health. In: Chaiyabutr N (ed) Milk production, IntechOpen, London, UK Krehl WA (1981) Discovery of the effect of tryptophan on niacin deficiency. Fed Proc 40:1527–1530

Kumazawa T, Kurihara K (1990a) Large enhancement of canine taste responses to sugars by salts. J Gen Physiol 95:1007–1018

Kumazawa T, Kurihara K (1990b) Large synergism between monosodium glutamate and 59– nucleotides in canine taste nerve responses. Am J Physiol 259:R420–R426

Kuo ZY (1967) The dynamics of behaviour development: an epigenetic view. Random House, New York

Ladd M, Raisz LG (1949) Response of the normal dog to dietary sodium chloride. Am J Physiol 159:149–152

Laflamme D (2008a) Effect of diet on loss and preservation of lean body mass in aging dogs and cats. Companion Animal Nutrition Summit, May 3–5, 2008, Charleston, South Carolina. pp 41–46

Laflamme DP (2008b) Pet food safety: dietary protein. Top Companion Anim Med 23:154–157

Laflamme DP, Hannah SS (2010) Pet dogs and vats. In: Pond KR (ed) Introduction to animal science (pond WG. John Wiley & Sons, New York, pp 526–552

Laflamme DP, Hannah SS (2013) Discrepancy between use of lean body mass or nitrogen balance to determine protein requirements for adult cats. J Feline Med Surg 15:691–697

Laflamme D, Danièlle G-M (2014) Nutrition of aging cats. Vet Clin North Am Sm Anim Pract 44:761–774

Lambie JG, Pezzali JG, Richards TL, Ellis JL, Verbrugghe A, Shoveller AK (2024) Phenylalanine requirements using the direct amino acid oxidation technique, and the effects of dietary phenylalanine on food intake, gastric emptying, and macronutrient metabolism in adult cats. J Anim Sci 102:skae009

Landsberg G, Milgram B, Mougeot I, Kelly S, de Rivera C (2017) Therapeutic effects of an alphacasozepine and L-tryptophan supplemented diet on fear and anxiety in the cat. J Feline Med Surg 19:594–602

Larsen TM, Toubro A, Astrup A (2003) Efficacy and safety of dietary supplements containing CLA for the treatment of obesity: evidence from animal and human studies. J Lipid Res 44:2234–2241

Larsen JA, Oberbauer AM (2023) Dog nutrition. In: Phillips CJC (ed) Encyclopedia of Animal Nutrition, 2nd edition. CABI, Wallingford, pp 176–179

Latimer HB (1938) The weights of the brain and of its parts, of the spinal cord and of the eyeballs in the adult cat J Comp Neurol 68:395–404

Latimer HB (1967) Variability in body and organ weights in the newborn dog and cat compared with that in the adult. Anat Red 157:449–456

Lawler DF, Larson BT, Ballam JM, Smith GK, Biery DN, Evans RH, Greeley EH, Segre M, Stowe HD, Kealy RD (2008) Diet restriction and ageing in the dog: major observations over two decades. Br J Nutr 99:793–805

Legrand-Defretin V (1994) Differences between cats and dogs: a nutritional view. Proc Nutr Soc 53:15–24

Lemaitre RN, King IB, Sotoodehnia N, Rea TD, Raghunathan TE, Rice KM, Lumley TS, Knopp RH, Cobb LA, Copass MK, Siscovick DS (2009) Red blood cell membrane alpha-linolenic acid and the risk of sudden cardiac arrest. Metabolism 58:534–540

Leung YB, Cave NJ, Heiser A, Edwards PJB, Godfrey AJR, Wester T (2020) Metabolic and immunological effects of intermittent fasting on a ketogenic diet containing medium-chain triglycerides in healthy dogs. Front Vet Sci 6:480

Levillain O, Parvy P, Hus-Citharel A (1996) Arginine metabolism in cat kidney. J Physiol 491:471– 477

Li P, Wu G (2018) Roles of dietary glycine, proline and hydroxyproline in collagen synthesis and animal growth. Amino Acids 50:29–38

Li P, Wu G (2020) Composition of amino acids and related nitrogenous nutrients in feedstuffs for animal diets. Amino Acids 52:523–542

Li P, Wu G (2022) Functional molecules of intestinal mucosal products in animal nutrition and health. Adv Exp Med Biol 1354:263–277

Li P, Wu G (2023a) Amino acid nutrition and metabolism in domestic cats and dogs. J Anim Sci Biotechnol 14:19

Li P, Wu G (2023b) Composition of nutrients in rendered animal-sourced feedstuffs. In: Phillips C (ed) Encyclopedia of animal nutrition, 2nd edn. CABI, Wallingford, Oxon, UK, pp 576–579

Li X, Li W, Wang H, Cao J, Maehashi K, Huang L, Bachmanov AA, Reed DR, LegrandDefretin V, Beauchamp GK (2005) Pseudogenization of a sweet-receptor gene accounts for cats’ indifference toward sugar. PLoS Genet 1:e27-35

Li X, Li W, Wang H, Bayley DL, Cao J, Reed DR, Bachmanov AA, Huang L, Legrand-Defretin V, Beauchamp GK, Brand JG (2006) Cats lack a sweet taste receptor. J Nutr 136(Suppl 7):1932S1934S

Li XL, Zheng SX, Jia SC, Song F, Zhou CP, Wu G (2020) Oxidation of energy substrates in tissues of largemouth bass (Micropterus salmoides). Amino Acids 52:1017–1032

Li P, He WL, Wu G (2021a) Composition of amino acids in foodstuffs for humans and animals. Adv Exp Med Biol 1332:189–210

Li XY, Zheng SX, Wu G (2021b) Nutrition and functions of amino acids in fish. Adv Exp Med Biol 1285:133–168

Li XY, Han T, Zheng SX, Wu G (2021c) Nutrition and functions of amino acids in aquatic crustaceans. Adv Exp Med Biol 1285:169–197

Li P, He WL, Wu G (2021d) Composition of amino acids in foodstuffs for humans and animals. Adv Exp Med Biol 1332:189–209

Li XY, He WL, Wu G (2023) Dietary glycine supplementation enhances the growth performance of hybrid striped bass (Morone saxatilis ~× Morone chrysops |) fed soybean meal-based diets. J Anim Sci 101:skad345

Lichtenberger M (2021) Thiaminase and its role in predatory pet fish (and other piscivores) nutrition. http://www.wetwebmedia.com/ca/volume_6/volume_6_1/thiaminase.htm: Accessed 19 Sept 2021

Loftus JP, Yazwinski M, Milizio JG, Wakshlag JJ (2014) Energy requirements for racing endurance sled dogs. J Nutr Sci 3:e34

Loncar D, Afzelius BA (1989) Ontogenetical changes in adipose tissue of the cat: convertible adipose tissue. J Ultrastruct Mol Struct Res 102:9–23

Macdonald DW, Apps PJ (1978) The social behaviour of a group of semi-dependent farm cats, Felis catus: a progress report. Carnivore Genet Newslett 3:256–268

MacDonald ML, Rogers QR, Morris JG (1985) Aversion of the cat to dietary medium-chain triglycerides and caprylic acid. Physiol Behav 35:371–375

Mansilla WD, Gorman A, Fortener L, Shoveller AK (2018) Dietary phenylalanine requirements are similar in small, medium, and large breed adult dogs using the direct amino acid oxidation technique. J Anim Sci 96:3112–3120

Mansilla WD, Templeman JR, Fortener L, Shoveller AK (2020a) Minimum dietary methionine requirements in Miniature Dachshund, Beagle, and Labrador Retriever adult dogs using the indicator amino acid oxidation technique. J Anim Sci 98:skaa324

Mansilla WD, Fortener L, Templeman JR, Shoveller AK (2020b) Adult dogs of different breed sizes have similar threonine requirements as determined by the indicator amino acid oxidation technique. J Anim Sci 98:skaa066

McCauley SR, Clark SD, Quest BW, Streeter RM, Oxford EM (2020) Review of canine dilated cardiomyopathy in the wake of diet-associated concerns. J Anim Sci 98:skaa155

McCue MD (2010) Starvation physiology: reviewing the different strategies animals use to survive a common challenge. Comp Biochem Physiol A 156:1–18

McDowell LR (1989) Vitamin C: vitamins in animal nutrition. San Diego, Academic Press, pp 365–387

McGeachin RL, Akin JR (1979) Amylase levels in the tissues and body fluids of the domestic cat (Felis catus). Comp Biochem Physiol B 63:437–439

Middleton RP, Lacroix S, Scott-Boyer MP, Dordevic N, Kennedy AD, Slusky AR, Carayol J, Petzinger-Germain C, Beloshapka A, Kaput J (2017) Metabolic differences between dogs of different body sizes. J Nutr Metab 2017:4535710

Mitchell WK, Phillips BE, Williams JP, Rankin D, Lund JN, Wilkinson DJ, Smith K, Atherton PJ (2015) The impact of delivery profile of essential amino acids upon skeletal muscle protein synthesis in older men. Clinical efficacy of pulse versus bolus supply. Am J Physiol 309:E450- 457

Miyazaki M, Yamashita T, Taira H, Suzuki A (2008) The biological function of cauxin, a major urinary protein of the domestic cat. In: Hurst JL, Beynon RJ, Roberts SC, Wyatt TD (eds) Chemical signals in vertebrates, vol 11. Springer, New York, pp 51–60

Montserrat-Malagarriga M, Castillejos L, Salas-Mani A, Torre C, Martín-Orúe SM (2024) The impact of fiber source on digestive function, fecal microbiota, and immune response in adult dogs. Animals 14:196

Morris JG (1999) Ineffective vitamin D synthesis in cats is reversed by an inhibitor of 7-dehydrocholestrol-∆7-reductase. J Nutr 129:903–908

Morris JG, Rogers QR, Pacioretty LM (1990) Taurine: an essential dietary nutrient for cats. J Small Anim Pract 31:502–509

Morris JG (2002) Idiosyncratic nutrient requirements of cats appear to be diet-induced evolutionary adaptations. Nutr Res Rev 15:153–168

Morris JG, Rogers QR (1978) Arginine: an essential amino acid for the cat. J Nutr 108:1944–1953

Morris JG, Trudell J, Pencovic T (1977) Carbohydrate digestion by the domestic cat (Felis catus). Br J Nutr 37:365–373

Mugford RA (1977) External influences on the feeding of carnivores. In: Maller O (ed) The chemical senses and nutrition (Kare MR. Academic Press, New York, pp 25–50

Mugford RA, Thorne C (1980) Comparative studies of meal patterns in pet and laboratory housed dogs and cats. In: Anderson RS (ed) Nutrition of the dog and cat. Pergamon Press, Oxford, UK, pp 3–14

National Research Council (NRC (2006) Nutrient requirements of dogs and cats. National Academies Press, Washington DC

National Research Council (NRC (2012) Nutrient requirements of swine. National Academies Press, Washington DC

Nogueira JPDS, He F, Mangian HF, Oba PM, de Godoy MRC (2019) Dietary supplementation of a fiber-prebiotic and saccharin-eugenol blend in extruded diets fed to dogs. J Anim Sci 97:4519–4531

Oberbauer AM, Larsen JA (2021) Amino acids in dog nutrition and health. Adv Exp Med Biol 1285:199–216

Oliveira R, HaeseI D, Kill JL, Lima A, Malini PV, Thompson GR (2016) Palatability of cat food with sodium pyrophosphate and yeast extract. Ciência Rural 46:2202–2205

Pawlosky R, Barnes A, Salem N Jr (1994) Essential fatty acid metabolism in the feline: relationship between liver and brain production of long-chain polyunsaturated fatty acids. J Lipid Res 35:2032–2040

Pekel AY, Mülazımo ˘glu SB, Acar N (2020) Taste preferences and diet palatability in cats. J Appl Anim Res 48:281–292

Pion PD, Kittleson MD, Rogers QR, Morris JG (1987) Myocardial failure in cats associated with low plasma taurine: a reversible cardiomyopathy. Science 237:764–768

Pion PD, Kittleson MD, Thomas WP, Skiles ML, Rogers QR (1992) Clinical findings in cats with dilated cardiomyopathy and relationship of findings to taurine deficiency. J Am Vet Med Assoc 201:267–274

Prosser CG (2021) Compositional and functional characteristics of goat milk and relevance as a base for infant formula. J Food Sci 86:257–265

Rabin B, Nicolosi RJ, Hayes KC (1976) Dietary influence on bile acid conjugation in the cat. J Nutr 106:1241–1246

Randall W, Johnson RF, Randall S, Cunningham JT (1985) Circadian rhythms in food intake and activity in domestic cats. Behav Neurosci 99:1162–1175

Ranz D, Gutbrod F, Corinna Eule C, Kienzle E (2002) Nutritional lens opacities in two litters of newfoundland dogs. J Nutr 132:1688S-1689S

Rassin DK, Sturman JA, Gaull GE (1978) Taurine and other free amino acids in milk of man and other mammals. Early Hum Dev 2:l–13

Reiland S (1978) Growth and skeletal development of the pig. Acta Radiol 358(Suppl):15–22

Rezaei R, Wu G (2022) Branched-chain amino acids regulate intracellular protein turnover in porcine mammary epithelial cells. Amino Acids 54:1491–1504

Rezaei R, Wu ZL, Hou YQ, Bazer FW, Wu G (2016) Amino acids and mammary gland development: nutritional implications for neonatal growth. J Anim Sci Biotechnol 7:20

Riond JL, Stiefel M, Wenk C, Wanner M (2003) Nutrition studies on protein and energy in domestic cats. J Anim Physiol Anim Nutr (berl) 87:221–228

Rivera NLM, Félix AP, Ferreira FM, da Silva AVF, Maiorka A (2011) Body measurements and serum lipid profile of overweight adult dogs fed diet with containing conjugated linoleic acid. Ciência Rural 41:2020–2025

Rivers JPW, Sinclair AJ, Crawford MA (1975) Inability of the cat to desaturate essential fatty acids. Nature 258:171–173

Robertson WG, Jones JS, Heaton MA, Stevenson AE, Markwell PJ (2002) Predicting the crystallization potential of urine from cats and dogs with respect to calcium oxalate and magnesium ammonium phosphate (struvite). J Nutr 132:1637S-1641S

Rodrıguez EMR, Alaejos MS, Romero CD (2001) Mineral concentrations in cow’s milk from the canary Island. J Food Compost Anal 14:419–430

Rogers QR, Morris JG (1979) Essentiality of amino acids for the growing kitten. J Nutr 109:718–723

Rogers QR, Phang JM (1985) Deficiency of pyrroline-5-carboxylate synthase in the intestinal mucosa of the cat. J Nutr 115:146–150

Rogers QR, Taylor TP, Morris JG (1998) Optimizing dietary amino acid patterns at various levels of crude protein for cats. J Nutr 128:2577S-2580S

Rogers QR, Morris JG, Freedland RA (1977) Lack of hepatic enzymatic adaptation to low and high levels of dietary protein in the adult cat. Enzyme 22:348–356

Rogers QR, Wigle AR, Laufer A, Castellanos VH, Morris JG (2004) Cats select for adequate methionine but not threonine. J Nutr 134:2046S-2049S

Romsos DR, Ferguson D (1983) Regulation of protein intake in adult dogs. J Am Vet Med Assoc 182:41–43

Romsos DR, Palmer HJ, Muiruri KL, Bennink MR (1981) Influence of a low carbohydrate diet on performance of pregnant and lactating dogs. J Nutr 111:678–689

Russell K, Murgatroyd PR, Batt RM (2002) Net protein oxidation is adapted to dietary protein intake in domestic cats (Felis silvestris catus). J Nutr 132:456–460

Rutherfurd-Markwick KJ, Rogers QR, Hendriks WH (2005) Mammalian isovalthine metabolism. J Anim Physiol Anim Nutr 89:1–10

Sawaya WN, Khalil JK, A F Al-Shalhat AF (1984) Mineral and vitamin content of goat’s milk. J Am Diet Assoc 84:433–435

Schaer M (1989) General principles of fluid therapy in small animal medicine. Vet Clin N Am Sm Anim Pract 19:203–213

Schoenherr W, Jewell J (1999) Effect of conjugated linoleic acid on body composition of mature obese Beagles. FASEB J 13:A262

Schneeman BO (1994) Carbohydrates: significance for energy balance and gastrointestinal function. J Nutr 124(Suppl 9):1747S-1753S

Schoenmakers I, Hazewinkel HAW, van den Brom WE (1999) Excessive Ca and P intake during early maturation in dogs alters Ca and P balance without long-term effects after dietary normalization. J Nutr 129:1068–1074

Schweigert FJ (1998) Metabolism of carotenoids in mammals. In: Liaaen-Jensen S, Pfander H (eds) Carotenoids (Britton G. Birkhäuser Verlag, Boston, pp 249–284

Schweigert FJ, Raila J, Wichert B, Kienzle E (2002) Cats absorb β-carotene, but it is not converted to vitamin A. J Nutr 132:1610S-1612S

Silvio J, Harmon DL, Gross KL, McLeod KR (2000) Influence of fiber fermentability on nutrient digestion in the dog. Nutrition 16:289–295

Sinclair AJ, Slattery W, McLean JG, Monger EA (1981) Essential fatty acid deficiency and evidence for arachidonate synthesis in the cat. Br J Nutr 46:93–96

Singh P, Banton S, Bosch G, Hendriks WH, Shoveller AK (2024) Beyond the bowl: understanding amino acid requirements and digestibility to improve protein quality metrics for dog and cat foods. Adv Exp Med Biol 1446:99–134. https://doi.org/10.1007/978-3-031-54192-6_5

Sparkes AH, Cannon M, Church D, Fleeman L, Harvey A, Hoenig M, Peterson ME, Reusch CE, Taylor S, Rosenberg D, ISFM (2015) ISFM consensus guidelines on the practical management of diabetes mellitus in cats. J Feline Med Surg 17:235–250

Summers SC, Stockman J, Larsen JA, Zhang L, Rodriguez AS (2020) Evaluation of phosphorus, calcium, and magnesium content in commercially available foods formulated for healthy cats. J Vet Intern Med 34:266–273

Supplee GC, Bellis B (1922) The coper content of cow’s milk. J Dairy Sci 5:455–467

Tôrres CL, Hickenbottom SJ, Quinton R. Rogers QR (2003) Palatability affects the percentage of metabolizable energy as protein selected by adult beagles. J Nutr 133:3516–3522

Trevizan L, de Mello KA, Bigley KE, Anderson WH, Waldron MK, Bauer JE (2010) Effects of dietary medium-chain triglycerides on plasma lipids and lipoprotein distribution and food aversion in cats. Am J Vet Res 71:435–440

Ugawa T, Kurihara K (1993) Large enhancement of canine taste responses to amino acids by salts. Am J Physiol 264:R1071-1076

Van Kruiningen HJ, Wojan LD, Stake PE, Lord PF (1987) The influence of diet and feeding frequency on gastric function in the dog. J Am Anim Hosp Assoc 23:146–153

Vendramini THA, Amaral AR, Pedrinelli V, Zafalon RVA, Rodrigues RBA, Brunetto MA (2020) Neutering in dogs and cats: current scientific evidence and importance of adequate nutritional management. Nutr Res Rev 33:134–144

Verbrugghe A, Bakovic M (2013) Peculiarities of one-carbon metabolism in the strict carnivorous cat and the role in feline hepatic lipidosis. Nutrients 5:2811–2835

Verbrugghe A, Hesta M (2017) Cats and carbohydrates: the carnivore fantasy? Vet Sci 4:55

Vhile SG, Skrede A, Ahlstrøm Ø, Hove K (2005) Comparative apparent total tract digestibility of major nutrients and amino acids in dogs (Canis familiaris), blue foxes (Alopex lagopus) and mink (Mustela vison). Anim Sci 81:141–148

Vinay P, Lemieux G, Gougoux A, Halperin M (1986) Regulation of glutamine metabolism in dog kidney in vivo. Kidney Int 29:68–79

Vuorinen A, Bailey-Hall E, Karagiannis A, Yu S, Roos F, Sylvester E, Wilson J, Irina Dahms I (2020) Safety of algal oil containing EPA and DHA in cats during gestation, lactation and growth. J Anim Physiol Anim Nutr (berl). 104:1509–1523

Wannemacher RW, McCoy JR (1966) Determination of optimal dietary protein requirements of young and old dogs. J Nutr 88:66–74

Washizu, T, Ishida T, Washizu M, Tomoda I, Kaneko JJ (1994) Changes in bile acid composition of serum and gallbladder bile in bile duct ligated dogs. J Vet Med Sci 56:299–303

Washizu T, Tanaka A, Sako T, Washizu M, Arai T (1999) Comparison of the activities of enzymes related to glycolysis and gluconeogenesis in the liver of dogs and cats. Res Vet Sci 67:205–206

Watford M (1985) Gluconeogenesis in the chicken: regulation of phosphoenolpyruvate carboxykinase gene expression. Fed Proc 44:2469–2074

Weber FL Jr, Maddrey WC, Walser M (1977) Amino acid metabolism of dog jejunum before and during absorption of keto analogues. Am J Physiol 232:E263-269

Weber FL Jr, Veach GL (1979) The importance of the small intestine in gut ammonium production in the fasting dog. Gastroenterology 77:235–240

Weber FL Jr, Friedman DW, Fresard KM (1988) Ammonia production from intraluminal amino acids in canine jejunum. Am J Physiol 254:G264-268

Weber MP, Stambouli F, Martin LJ, Dumon HJ, Biourge VC, Nguyen PG (2002) Influence of age and body size on gastrointestinal transit time of radioopaque markers in healthy dogs. Am J Vet Res 63:677–682

Weber MP, Martin LJ, Biourge VC, Nguyen PG, Dumon HJ (2003) Influence of age and body size on orocecal transit time as assessed by use of the sulfasalzine method in healthy dogs. Am J Vet Res 64:1105–1109

Wester TJ, Weidgraaf K, Hekman M, Ugarte CE, Forsyth SF, Tavendale MH (2015) Amino acid oxidation increases with dietary protein content in adult neutered male cats as measured using [1-13C]leucine and [15N2]urea. J Nutr 145:2471–2478

White TD, Boudreau JC (1975) Taste preferences of the cat for neurophysiologically active compounds. Physiol Psychol 3:405–410

Widdowson EM (1985) Development of the digestive system: comparative animal studies. Am J Clin Nutr 41(Suppl):384–390

Wildgrube HJ, Stockhausen H, Petri J, Füssel U, Lauer H (1986) Naturally occurring conjugated bile acids, measured by highperformance liquid chromatography, in human, dog, and rabbit bile. J Chromatogr 353:207–213

Williams CC, Cummins KA, Hayek MG, Davenport GM (2001) Effects of dietary protein on whole-body protein turnover and endocrine function in young-adult and aging dogs. J Anim Sci 79:3128–3136

Wolf LGG, Mehlman MA (1972) Subcellular distribution of pyruvate carboxylase and phosphoenolpyruvate carboxykinase in dog liver and kidney. Proc Soc Exp Biol Med 141:532–535

Woods JE, Besch EL (1971) Dissipation of body heat in dogs in a controlled environment. Physiologist 14:252

Wu G (1998) Amino acid metabolism in the small intestine. Trends Comp Biochem Physiol 4:39–74

Wu G (2009) Amino acids: metabolism, functions, and nutrition. Amino Acids 37:1–17 Wu G (2016) Dietary protein intake and human health. Food Funct 7:1251–1265

Wu G (2018) Principles of animal nutrition. CRC Press, Boca Raton, Florida

Wu G (2020a) Metabolism and functions of amino acids in sense organs. Adv Exp Med Biol 1265:201–217

Wu G (2020b) Important roles of dietary taurine, creatine, carnosine, anserine and 4-hydroxyproline in human nutrition and health. Amino Acids 52:329–360

Wu G (2022) Amino acids: biochemistry and nutrition, 2nd edn. CRC Press, Boca Raton, Florida

Wu G, Morris SM Jr (1998) Arginine metabolism: nitric oxide and beyond. Biochem J 336:1–17

Wu G, Li P (2022) The “ideal protein” concept is not ideal in animal nutrition. Exp Biol Med 247:1191–1201

Wu G, Borbolla AG, Knabe DA (1994) The uptake of glutamine and release of arginine, citrulline and proline by the small intestine of developing pigs. J Nutr 124:2437–2444

Wu G, Bazer FW, Satterfield MC, Li XL, Wang XQ, Johnson GA, Burghardt RC, Dai ZL, Wang JJ, Wu ZL (2013) Impacts of arginine nutrition on embryonic and fetal development in mammals. Amino Acids 45:241–256

Wu G, Cross HR, Gehring KB, Savell JW, Arnold AN, McNeill SH (2016a) Composition of free and peptide-bound amino acids in beef chuck, loin, and round cuts. J Anim Sci 94:2603–2613

Wu ZL, Hou YQ, Hu SD, Bazer FW, Meininger CJ, McNeal CJ, Wu G (2016b) Catabolism and safety of supplemental L-arginine in animals. Amino Acids 48:1541–1552

Xenoulis PG, Palculict B, Allenspach K, Steiner JM, Van House AM, Suchodolski JS (2008) Molecular-phylogenetic characterization of microbial communities imbalances in the small intestine of dogs with inflammatory bowel disease. FEMS Microbiol Ecol 66:579–589

Yu YM, Wagner DA, Tredget EE, Walaszewski JA, Burke JF, Young VR (1990) Quantitative role of splanchnic region in leucine metabolism: L-[1-13C,15N]leucine and substrate balance studies. Am J Physiol 259:E36-51

Yu YM, Burke JF, Tompkins RG, Martin R, Young VR (1996) Quantitative aspects of interorgan relationships among arginine and citrulline metabolism. Am J Physiol 271:E1098-1109

Yu S, Rogers QR, Morris JG (1997) Absence of a salt (NaCl) preference or appetite in sodium-replete or depleted kittens. Appetite 29:1–10

Yu S, Rogers QR, Morris JG (2001) Effect of low levels of dietary tyrosine on the hair colour of cats. J Small Anim Pract 42:176–180

Zafalon RVA, Risolia LW, Pedrinelli V, Vendramini THA, Rodrigues RBA, Amaral AR, Kogika MM, Brunetto MA (2020a) Vitamin D metabolism in dogs and cats and its relation to diseases not associated with bone metabolism. J Anim Physiol Anim Nutr 104:322–342

Zafalon RVA, Risolia LW, Vendramini THA, Ayres Rodrigues RB, Pedrinelli V, Teixeira FA et al (2020b) Nutritional inadequacies in commercial vegan foods for dogs and cats. PLoS ONE 15:e0227046

Zaghini G, Biagi G (2005) Nutritional peculiarities and diet palatability in the cat. Vet Res Commun 29(Suppl 2):39–44

Zhai H, Adeola O (2011) Apparent and standardized ileal digestibilities of amino acids for pigs fed corn-and soybean meal-based diets at varying crude protein levels. J Anim Sci 89:3626–3633

Zhang Q, Hou YQ, Bazer FW, He WL, Posey EA, Wu G (2021) Amino acids in swine nutrition and production. Adv Exp Med Biol 1285:81–107

Zoran DL (2010) Obesity in dogs and cats: a metabolic and endocrine disorder. Vet Clin North Am Small Anim Pract 40:221–239

Zoran DL (2023) Cat (Felis catus), essential nutrients. In: Phillips CJC (ed) Encyclopedia of Animal Nutrition, 2nd edition. CABI, Wallingford, pp 100–102

Related topics:
Authors:
Guoyao Wu
Texas A&M University
Recommend
Comment
Share
Guoyao Wu
Texas A&M University
27 de abril de 2024
A new book "Nutrition and Metabolism of Dogs and Cats" was recently published by Springer in the well-established series of "Advances in Experimental Medicine and Biology". Here is the link to this book: https://link.springer.com/book/10.1007/978-3-031-54192-6
Recommend
Reply
Profile picture
Would you like to discuss another topic? Create a new post to engage with experts in the community.
Join Engormix and be part of the largest agribusiness social network in the world.