draft · 2018. 2. 12. · 32 review explores behavior of the tetracycline class of antibiotics from...

54
Draft Fate of Tetracycline Antibiotics in Dairy Manure-Amended Soils Journal: Environmental Reviews Manuscript ID er-2017-0041.R1 Manuscript Type: Review Date Submitted by the Author: 22-Sep-2017 Complete List of Authors: Pollard, Anne; Washington State University, Department of Crop and Soil Sciences Morra, Matthew; University of Idaho, Department of Soil & Water Systems Keyword: Antibiotic, Tetracycline, Soil, Dairy, Manure https://mc06.manuscriptcentral.com/er-pubs Environmental Reviews

Upload: others

Post on 20-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Draft

    Fate of Tetracycline Antibiotics in Dairy Manure-Amended

    Soils

    Journal: Environmental Reviews

    Manuscript ID er-2017-0041.R1

    Manuscript Type: Review

    Date Submitted by the Author: 22-Sep-2017

    Complete List of Authors: Pollard, Anne; Washington State University, Department of Crop and Soil Sciences Morra, Matthew; University of Idaho, Department of Soil & Water Systems

    Keyword: Antibiotic, Tetracycline, Soil, Dairy, Manure

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    1

    Fate of Tetracycline Antibiotics in Dairy Manure-Amended Soils 1

    2

    Anne T. Pollard1 and Matthew J. Morra

    a. Department of Soil & Water Systems, University of 3

    Idaho, 875 Perimeter Drive MS 2340, Moscow, ID 83844-2340, USA 4

    5

    a Department of Soil & Water Systems, University of Idaho, 875 Perimeter Drive MS 2340, 6

    Moscow, ID 83844-2340, USA. [email protected] 7

    8

    Corresponding author: 9

    Anne T. Pollard 10

    Department of Crop and Soil Sciences, Washington State University, Johnson Hall 233E, PO 11

    Box 646420, Pullman, WA 99164-6420, USA 12

    Phone: 208-301-8469 13

    Email: [email protected] 14

    15

    16

    17

    Word count: 12,176 (including references and figures) 18

    1 Present address: Department of Crop and Soil Sciences, Washington State University, Johnson

    Hall 233E, PO Box 646420, Pullman, WA 99164-6420, USA. [email protected]

    Page 1 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    2

    Abstract: The U.S. dairy industry has changed significantly during the past 20 years. The 19

    number of dairies declined 63% from 1997 to 2012 due to the rise in concentrated animal feeding 20

    operations (CAFOs) and the concomitant decline of small dairy farms. Efficient and cost-21

    effective dairies adhering to the CAFO business design are praised for their high milk 22

    production. However, with a per capita daily manure production of 55 kg, storage and disposal of 23

    manure at these large operations pose significant management challenges and environmental 24

    risks. Application to surrounding agricultural fields is a common practice for disposing of 25

    manure, but the fate and consequences of antibiotics present in dairy waste are issues of great 26

    concern. Although antibiotics in the environment promote microbial resistance, their risks to 27

    humans and the environment are not completely known. Understanding and predicting the fate of 28

    antibiotics from dairy manure in soils is complicated by the variability and complex interactions 29

    of soil factors in addition to the diversity of chemicals of emerging concern (CECs), their 30

    amphoteric structures, and potential antagonistic and synergistic interactions among CECs. This 31

    review explores behavior of the tetracycline class of antibiotics from dairy manure in the soil 32

    environment. Tetracycline fate in soils depends significantly on soil pH, ionic strength, and soil 33

    organic matter. Molecular charge and physicochemical properties of tetracyclines at typical soil 34

    pHs encourage strong sorption to soils; however, this interaction is complicated by organic 35

    matter and metals, and may also encourage development of antibiotic resistance. Furthermore, 36

    tetracycline degradation products exhibit distinct properties from their parent compounds that 37

    also must be considered. Increased knowledge of the behavior of tetracycline antibiotics in soil is 38

    needed to enable mitigation of their potential risks. 39

    Key words: Dairy manure; Antibiotic; Tetracycline; Soil 40

    41

    Page 2 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    3

    1. Introduction 42

    The transition from smallholder dairies to CAFO-style dairy operations has resulted in 43

    massive generation of animal waste in a physically restricted location. While spreading manure 44

    on surrounding farmland is a centuries-old practice of waste removal, the extensive amount of 45

    waste that is being applied near CAFO dairies has triggered mounting concerns about the 46

    potential human and environmental risks posed by CECs (chemicals of emerging concern) in 47

    manure. Veterinary pharmaceuticals in concentrations of micrograms to nanograms per liter have 48

    been found in surface and groundwaters located proximate to manure-applied soils, illustrating 49

    the mobility of these CECs in soil and therefore their potential threat to the environment (Daghrir 50

    and Drogui 2013). Given their low cost, minimal adverse side effects, and broad spectrum 51

    application, compounds within the tetracycline family are among the most highly used 52

    antibiotics for livestock and humans (Chopra and Roberts 2001; USDA 2008a; USFDA 2015). 53

    Daghrir and Drogui (2013) reviewed tetracycline occurrence and toxicity, with an emphasis on 54

    mitigating water treatment processes. Additional reviews exist on the presence and behaviors of 55

    veterinary pharmaceuticals in soils (Tolls 2001; Thiele-Bruhn 2003); however, the fate of 56

    tetracycline antibiotics in soils following dairy manure application has not been comprehensively 57

    assessed. The variety of chemical compounds in dairy manure, the diversity of soil types to 58

    which dairy waste is applied, and the various waste application practices complicate such 59

    evaluation. This review addresses the occurrence, properties, and fate, including sorption, 60

    transport, and degradation, of tetracycline (TC), chlortetracycline (CTC), and oxytetracycline 61

    (OTC) antibiotics in dairy manure-amended soils. 62

    2. Antibiotics in the dairy industry 63

    Page 3 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    4

    The source of detected pharmaceuticals in surface and groundwater has historically been 64

    attributed to human use and the consequent release of effluents from sewage treatment plants 65

    (Kolpin et al. 2002). However, the dominant consumer of pharmaceuticals in the United States is 66

    livestock, encompassing swine, poultry, beef, and dairy cattle (USEPA 2013). Antibiotics are 67

    used in the dairy industry to treat existing disease, to prevent disease, and for growth promotion, 68

    though the latter is used only in pre-lactating calves (USDA 2008b; Chee-Sanford et al. 2009; 69

    USEPA 2013). The use of antibiotics for growth promotion in livestock is banned in the 70

    European Union, but remains a common practice in the USA and China, albeit not frequently for 71

    dairy cattle (Sarmah et al. 2006; USEPA 2013; Zhou et al. 2013). Tetracycline antibiotics are 72

    used in dairy operations to treat respiratory infections, diarrhea, navel infections, reproductive 73

    disorders, mastitis, lameness, and other ailments (USDA 2009). In 2007, TC was given to 25% 74

    of heifers to treat respiratory infections, down from 34.3% in 2002; 55% of heifers to treat 75

    diarrhea and digestive disorders, a significant increase from 11.8% in 2002; and 42% of heifers 76

    to treat lameness, unchanged from 2002 (USDA 2008a, 2009). 77

    Antibiotics are administered to dairy cattle orally through feed and water, by injection, 78

    topically, or via intramammary or intrauterine infusion (Zwald et al. 2004; Zhou et al. 2013). In a 79

    survey of 99 conventional dairies from the states of New York, Wisconsin, Michigan, and 80

    Minnesota, 98% reportedly used intramammary antibiotic infusions prophylactically in non-81

    lactating cows (Zwald et al. 2004). In 2007, 49.5% of all US dairy operations fed milk replacer 82

    medicated with a combination of OTC and neomycin to preweaned heifers (USDA 2008a). 83

    Additionally, a survey of 80% of all US dairies indicated that 82% of all cows were given 84

    antibiotics at dry-off, or the period between milk stasis and subsequent calving (USDA 2008a). 85

    These results are similar to a survey of 201 dairies in The Netherlands in which 82.8% were 86

    Page 4 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    5

    found to administer antibiotics on dry cows (Barkema et al. 1998). From 2002 to 2007, the 87

    percent of US dairy operation that add CTC or OTC to weaned heifer rations decreased from 88

    62.4% to 14.4% for CTC and from 21.5% to 10.9% for OTC (USDA 2009). 89

    It is difficult to accurately assess antibiotic use in dairies because 1) operations are not 90

    required to report data; 2) data on which studies have been based result from voluntary 91

    participation, which introduces a bias since non-responder dairy operations may differ from the 92

    responders; 3) the majority of studies group beef and dairy cattle into a single category; 4) 93

    standardized record keeping systems do not exist; 5) most respondents rely on memory regarding 94

    antibiotic use, frequency, duration, and prevalence; 6) it is not possible to verify the accuracy of 95

    reported findings; and 7) extra-label drug usage occurs, meaning antibiotics are used on animals 96

    or for conditions that the drug is not specified (Zwald et al. 2004; Van Boeckel et al. 2015). For 97

    example, 55% of conventional dairies surveyed in Wisconsin indicated extra-label use of 98

    intramammary antibiotics (Pol and Ruegg 2007). An estimated 15.4 million kg of antimicrobials 99

    were sold and distributed in 2014 specifically for domestic food-producing animals with 100

    tetracyclines accounting for the greatest share overall at 43% (USFDA 2015). Furthermore, from 101

    2009 to 2014, sales and distribution of tetracyclines for domestic food-producing animals 102

    increased 25% (USFDA 2015). However, the report does not specify the amount distributed per 103

    livestock species. Antibiotics are poorly metabolized in vivo and estimates of 30-90% of the 104

    parent compound and its active metabolites are excreted in animal urine and feces (Chee-Sanford 105

    et al. 2009; Daghrir and Drogui 2013). Moreover, because antibiotics are given for prophylactic, 106

    metaphylactic, and growth-promoting purposes, the antibiotic load present in manure far 107

    surpasses historic quantities when more conservative practices were employed. Through land-108

    application of antibiotic-rich animal waste in concentrated geographical regions, the soil serves 109

    Page 5 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    6

    as a repository for pharmaceuticals from which tetracyclines may leach into groundwater, enter 110

    surface waters via overland flow, or persist in soil (Kwon 2011; Zhou et al. 2013; Pan and Chu 111

    2017). 112

    Although a thorough investigation of the association between land-application of dairy 113

    manure and development of antibiotic resistance in soil bacteria is beyond the scope of this 114

    review, its great importance warrants a brief explanation. In addition to large quantities of 115

    residual antibiotics in animal manure, commensal bacteria present in animal guts are likewise 116

    present in high concentrations in animal waste (Chee-Sanford et al. 2009). Due to frequent 117

    ingestion of antibiotics, commensal bacteria contain a high concentration of antibiotic-resistant 118

    genes (ARG), which confer resistance to antibiotics in microorganisms (Wichmann et al. 2014). 119

    In their metagenomic analysis of dairy cow manure to assess ARG present in dairy cow gut 120

    bacteria, Wichmann et al. (2014) identified 80 unique ARG, including three that confer 121

    tetracycline resistance: tetW, tetO, and an unspecified gene. Chambers et al. (2015) likewise 122

    conducted a similar metagenomics analysis of ARG in dairy cow manure and 75% of detected 123

    ARG were of the tetracycline class. Antibiotics present in animal waste, in addition to excreted 124

    commensal bacteria, create a conducive environment for proliferation of ARG (Storteboom et al. 125

    2007; Kyselková et al. 2015a,b; Peng et al. 2016). Manure samples from antibiotic-treated dairy 126

    cows contained nearly the same quantity and distribution of ARG as manure samples from dairy 127

    cows with no previous antibiotic exposure, illustrating the ease with which ARG transfer to other 128

    bacteria by mobile genetic elements (Wichmann et al. 2014). In a survey of tetracycline resistant 129

    genes in cow waste, water, and soil samples at six grassland-based dairy farms, 14 unique 130

    tetracycline ARG were detected among the animal and environmental samples, with tetW and 131

    tetQ detected in nearly every sample at each of the farms (Santamaría et al. 2011). It appears that 132

    Page 6 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    7

    tetW and tetQ arrived in soil and water samples as a result of horizontal gene transfer since those 133

    ARG have the exact genetic sequences as the tetW and tetQ detected in animal waste samples. 134

    Though research frequently focuses on the introduction of pathogenic microbes to the 135

    environment from animal manure (Salgado et al. 2011), data indicate that manure application is 136

    directly associated with increasing numbers of antibiotic-resistant bacteria in soil (Sengeløv et al. 137

    2003; Peng et al. 2014). Residual antibiotics in land-applied manure exert a selective pressure on 138

    resident soil microbes, causing a shift in the microbial community toward more resistant 139

    microorganisms (Boxall et al. 2003; Pruden et al. 2006; Chessa et al. 2016). Recent studies show 140

    that very low antibiotic concentrations in the environment can induce selection for resistance 141

    (Gullberg et al. 2011; Liu et al. 2011). The soil environment may serve as a reservoir for ARG, 142

    thereby contributing to the inefficacy of antibiotics in humans, as many of the drugs commonly 143

    administered to livestock are also routinely used by humans (Storteboom et al. 2007). In addition 144

    to widespread concern for development of antibiotic resistance, some antibiotics have also 145

    demonstrated an ability to behave as hormones, thus having endocrine disrupting effects on 146

    organisms (Hamscher et al. 2005). 147

    2.1. Physical and chemical properties of tetracyclines 148

    The tetracycline class of antibiotics contains some of the most widely used broad-spectrum 149

    antibiotics for humans and animals because of their low cost and antimicrobial efficacy (Daghrir 150

    and Drogui 2013; Chang et al. 2015). Tetracyclines function as antibacterial agents by attaching 151

    to a transfer-RNA binding site on the 30S ribosomal subunit in bacteria, thereby inhibiting 152

    protein synthesis (Carter et al. 2000). Numerous tetracycline antibiotics are used in the livestock 153

    industry, with tetracycline (TC), chlortetracycline (CTC), and oxytetracycline (OTC) the most 154

    commonly used antibiotics worldwide (Granados-Chinchilla and Rodríguez 2017). Among three 155

    Page 7 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    8

    dairy farms sampled in China, CTC constituted 86% of total antibiotics excreted by cows, 156

    equating to 3.66 mg d-1

    cow-1

    (Zhou et al. 2013). In addition to therapeutic treatment of mastitis, 157

    foot infections, respiratory disease, and metritis, they are also used prophylactically for disease 158

    prevention in dry cows and for growth promotion in pre-lactating cows (Zwald et al. 2004; Pol 159

    and Ruegg 2007; USEPA 2013). Chlortetracycline and OTC are among the top ten antibiotics 160

    licensed for use as growth promoters in the United States (Hao et al. 2014). Tetracyclines are 161

    poorly absorbed in the digestive tract, with an estimated 25% excreted in feces and an additional 162

    50-60% excreted as the parent molecule or an active metabolite in urine (Feinman and Matheson 163

    1978; Boothe 2015). The concentration and frequency of tetracyclines reported in dairy manure 164

    vary. Oxytetracycline has been found in high concentrations in fresh dairy calf manure, 165

    illustrating its poor intestinal absorption (De Liguoro et al. 2003). From three dairy farms 166

    sampled in China, tetracyclines were present in fresh feces at only one farm, with 1450 µg kg-1

    167

    CTC and 17 µg kg-1

    TC (Zhou et al. 2013). 168

    Tetracyclines are large molecules with molecular weights ranging from 480.9 g mol-1

    for TC 169

    to 528 g mol-1

    for CTC. They share a naphthalene ring structure with three ionizable functional 170

    groups: tricarbonyl (pKa 3.2 to 3.3), phenolic β-diketone (7.46 to 7.78), and dimethylamine (8.9 171

    to 9.6) (Teixido et al. 2012) (Fig. 1). The multiple polar functional groups give the molecules 172

    amphoteric characteristics, whereby they may be a cation, zwitterion, or anion (Sassman and Lee 173

    2005) (Table 1). At soil pH values between approximately 4 and 8, tetracyclines will exist as 174

    zwitterions with a positive charge on the amine and a negative charge on the hydroxyl group in 175

    the tricarbonyl (Sassman and Lee 2005; Li et al. 2010). 176

    The distinctive physical structures of the tetracycline molecules attribute unique chemical 177

    properties to the individual antibiotics. Tetracycline displays the basic structure of the class, with 178

    Page 8 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    9

    hydrogen at C10 of the D-ring replaced by a chlorine atom to produce CTC, and a hydroxyl 179

    group at C6 of the B-ring substituted for the hydrogen to produce OTC (Teixido et al. 2012) 180

    (Fig. 1). In addition to the structures of the parent molecules, tetracyclines can undergo 181

    transformations to form numerous bioactive epimers (Kwon 2011). Due to substitutions at C10 182

    and C6, CTC and OTC are more reactive in the environment than TC (Loftin et al. 2008). CTC 183

    has greater solubility and mobility than TC or OTC (Kwon 2011; Teixido et al. 2012). In terms 184

    of degradability, CTC is the most easily degraded, followed by OTC, then TC (Loftin et al. 185

    2008). Increased reactivity, solubility, and degradability of CTC may be attributed to its reported 186

    instability at alkaline pH (Figueroa et al. 2004). Avisar et al. (2010) reported lower sorption of 187

    OTC to montmorillonite clay than TC as a result of molecular structure differences. The 188

    presence of a hydroxyl adjacent to the cationic amine residue on OTC can cause steric hindrance, 189

    decreasing its ability to sorb to clay surfaces. Tetracyclines are essentially non-volatile, as 190

    indicated by their low Henry’s constants (3.91 x 10-26

    to 3.45 x 10-24

    atm m3 mol

    -1) (Daghrir and 191

    Drogui 2013). Their high water solubilities (0.008-0.062 mol L-1

    ) and low log Kow values (-1.25 192

    to -0.62) suggest a hydrophilic nature (Daghrir and Drogui 2013). Furthermore, a negative log 193

    Kow suggests that any tetracycline partitioning to soils is a result of electrostatic interactions, ion 194

    exchange, and surface complexation (Jones et al. 2005). 195

    2.2. Occurrence of manure-borne tetracyclines in soils 196

    Predicting sorption, transport, and bioavailability of tetracycline antibiotics following land-197

    application of dairy manure is challenging. Relatively few field studies have characterized the 198

    fate of antibiotics in soil following land-application of manure, and even fewer specifically 199

    address dairy manure. The potential environmental risks from tetracycline introduction to the 200

    environment from manure application include development and spread of antibiotic resistance 201

    Page 9 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    10

    and introduction of antibiotic resistance to the food chain (Storteboom et al. 2007). Moreover, 202

    antibiotic-resistant organisms exhibit unique physiological structures and processes compared to 203

    their non-resistant counterparts (Walczak et al. 2011), resulting in unique interactions with the 204

    soil environment. For example, in packed sand column studies comparing movement of 205

    tetracycline-resistant (tetR) and tetracycline-susceptible (tet

    S) E. coli isolated from dairy manure, 206

    the tetR strains had greater mobility than the tet

    S. The lower zeta potentials of the tet

    R resulted in 207

    greater repulsive forces between the tetR and negatively charged sand surfaces, thus facilitating 208

    increased transport. The composition of outer membrane proteins between the resistant and 209

    susceptible strains was also very different, which may have contributed to the increased mobility 210

    of the tetR strains (Walczak et al. 2011). Drastic alteration of microbial physiology and behavior 211

    due to concentrated and prolonged exposure to environmental antibiotics will necessitate 212

    reevaluation of the current microbiological knowledge base. 213

    Due to their widespread use and high sorption capacity, tetracyclines have been detected in 214

    substantial concentrations in soils. In an analysis of antibiotics present in fresh cattle, swine, and 215

    chicken manure, the tetracycline class of antibiotics was detected most frequently (81-91%) and 216

    in the greatest concentrations, with CTC predominating at an average concentration of 8060 ± 217

    23080 µg kg-1

    (Hou et al. 2015). The concentration of CTC in cattle manure was far less than for 218

    swine or chicken manure, averaging approximately 300 µg kg-1

    . Moreover, the concentration of 219

    CTC and TC detected in adjacent manure-amended soils was very similar to the concentrations 220

    in the fresh manure. For example, the CTC concentration in fresh manure was approximately 221

    1550 µg kg-1

    while it was detected at ~ 1500 µg kg-1

    in the adjacent field (Hou et al. 2015). In a 222

    study from Germany of tetracycline behavior in a sandy soil (91.5% sand) repeatedly fertilized 223

    with liquid pig manure containing initial concentrations of TC and CTC of 14.1-41.2 mg kg-1

    and 224

    Page 10 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    11

    0.9-1.0 mg kg-1

    , respectively, the average concentrations detected in soil were 158 µg kg-1

    for TC 225

    and 8.3 µg kg-1

    for CTC (Hamscher et al. 2005). Hamscher et al. (2005) additionally found 226

    consistently high concentrations of tetracyclines in the top 0-30 cm of soil for 2 years, illustrating 227

    their persistence and consequent accumulation in soils with repeated manure application 228

    (Hamscher et al. 2002, 2005), as was similarly determined in more recent studies examining 229

    tetracycline occurrence in swine manure (Martinez-Carballo et al. 2007; Qiao et al. 2012). 230

    In a study from northern China comparing antibiotic concentrations in soils after either 231

    swine or dairy wastewater application (initial concentrations not specified), the levels of 232

    tetracyclines differed significantly between the animal operations. No tetracyclines were detected 233

    in the dairy field soil, but between 0 and 667 µg kg-1

    were detected in the swine fields (depth of 234

    soil sampling not specified) (Zhou et al. 2013). Tetracyclines were also not detected in manure 235

    lagoon samples or soil leachates (sampling details not provided) from two conventional and two 236

    organic German dairies, leading Kemper et al. (2008) to conclude that dairies are an insignificant 237

    source of antibiotics to the environment. However, their negative results may also be influenced 238

    by the tendency of tetracyclines to sorb strongly to soil particles and resist leaching. Contrasting 239

    results from different animal manure studies can be attributed to the amount of antibiotics 240

    originally administered; different animals’ physiologies and resultant metabolism of 241

    tetracyclines; number of animals contributing to manure output, thus drug concentrations in the 242

    manure; and overall different operational systems and methods utilized at swine and dairy farms. 243

    Moreover, studies show that composition and concentration of antibiotics in animal manure are 244

    affected by animal type, life stage, and husbandry practices (Hou et al. 2015). For example, the 245

    concentration of CTC in fresh piglet manure was far greater (~ 35000 µg kg-1

    ) than in sow 246

    manure (~14000 µg kg-1

    ) likely from high dosages of CTC administered to piglets for disease 247

    Page 11 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    12

    prevention (Hou et al. 2015). Furthermore, antibiotics at the dairies in Kemper et al.’s study 248

    (2008) were only given for treatment of disease and typically to individual dairy cows, whereas 249

    common practice on pig farms is to routinely administer antibiotics to entire herds for 250

    prophylactic and metaphylactic purposes (Zhou et al. 2013). Similarly, Zhou et al. (2013) stated 251

    that antibiotic use on dairy farms in China was primarily for treatment and prevention of disease, 252

    and not for growth promotion, a common purpose of antibiotic administration on swine farms. 253

    Quantities of antibiotics, including tetracyclines, in manure from a beef feedlot were 254

    significantly greater than levels in dairy manure (specific values were not reported), presumably 255

    resulting from high doses given to beef for growth promotion (Storteboom et al. 2007). These 256

    studies highlight the variability in reported concentrations of tetracyclines in animal waste as a 257

    result of diverse animal husbandry methods, such as administered amounts of drugs and 258

    operating procedures. 259

    2.3. Tetracycline sorption in soils 260

    Aside from minor differences in their chemical properties, sorption behavior of TC, OTC, 261

    and CTC on clay surfaces is extremely similar due to their structural resemblance (Figueroa et al. 262

    2004). The main determinants of tetracycline sorption to soils in order of significance are pH, 263

    which controls speciation and Coulombic or electrostatic interactions; CEC; metal content 264

    (especially Fe2O3, K2O and to a lesser extent Al2O3), which influences surface complexation 265

    with metal oxides and cation bridging; and soil texture, such as clay content (Jones et al. 2005; 266

    Sassman and Lee 2005; Teixido et al. 2012; Peng et al. 2014) (Fig. 2). Depending on the soil 267

    physicochemical properties, sorption coefficient Kd values for tetracyclines spiked in soils have 268

    been measured ranging from < 200 to over 5,000 L Kg-1

    (Table 2) (Rabølle and Spliid 2000; 269

    Sassman and Lee 2005; Teixido et al. 2012; Chessa et al. 2016). Through batch-equilibration 270

    Page 12 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    13

    sorption studies in which 30 unique soils devoid of historical animal usage were spiked with 271

    OTC, Jones et al. (2005) found that texture, oxide content, and cation exchange capacity 272

    significantly influence OTC sorption in soils with ≤ 4% organic carbon (Table 2). Pearson 273

    correlation and a principle components analysis correlating OTC sorption to 16 physicochemical 274

    properties of the 30 soils showed OTC sorption coefficients (Kd) most positively correlated to 275

    CEC, percent clay, surface area, Fe-oxide content, and exchangeable Mg2+

    , Ca2+

    , and Na

    +. 276

    Sorption was negatively correlated to percent sand. In general, tetracyclines sorb quickly and 277

    strongly to soils, especially at low pH and in soils with high clay content, as seen in the batch 278

    sorption studies of Sassman and Lee (2005) (Table 2) where they spiked variable soils with OTC 279

    concentrations comparable to those found at livestock operations. Sorption of tetracyclines onto 280

    ‘Toronto’ soil (pH 4.18) averaged approximately 30 times greater than sorption onto ‘Drummer’ 281

    soil (pH 7.49), though they had equal 21% clay contents. Chessa et al. (2016) drew similar 282

    conclusions from batch equilibration studies where alkaline and acid soils were 1) spiked with 283

    TC, or 2) incorporated with TC-spiked cattle manure (Table 2). The Kd value was five times 284

    greater in a pH 5.77 soil with 41% clay compared to a pH 7.6 soil with 16.6% clay (Chessa et al. 285

    2016). Moreover, the slight change in TC sorption after manure addition further suggests that 286

    clay content and pH, rather than organic matter content, strongly influence TC sorption to soils. 287

    A wide variability of Kd values among 30 soil samples from five represented soil orders 288

    indicates that OTC sorption to soils cannot be determined solely based on soil order (Jones et al 289

    2005). Nevertheless, Spodosol samples displayed the lowest attraction for OTC, which may be 290

    due to low surface area of the coarse-textured soils (Jones et al. 2005). Other studies have also 291

    determined that finer-textured soils display greater capacity for OTC sorption (Rabølle and 292

    Spliid 2000; Loke et al. 2002) (Table 2). In a batch-equilibration study analyzing OTC-spiked 293

    Page 13 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    14

    soils, between 95 – 99% of OTC was sorbed to loamy sand and sandy loam soils, with Kd values 294

    of 417 and 1026 mL g-1

    , respectively (Rabølle and Spliid 2000) (Table 2). Despite this range in 295

    Kd values, infrequent detection of tetracyclines in groundwater is consistent with their high Kd 296

    values. In a comparison of CTC sorption kinetics onto two soils with similar pH and organic 297

    matter content, but divergent particle size distribution, CTC rapidly sorbed onto both the sandy 298

    loam and heavy clay soil, further support that tetracyclines have low probability of leaching 299

    through the soil profile (Allaire et al. 2006). 300

    Sorption of tetracyclines to montmorillonite and kaolinite clays depends on solution pH, 301

    ionic strength, and organic matter content (Figueroa et al. 2004; Avisar et al. 2010; Zhao et al. 302

    2011 and 2015). Estimating tetracycline sorption in soils based on Kd values attained from 303

    research with pure clay minerals may be misleading, since clay surfaces in actual soils may be 304

    coated with organic matter or metal oxides, thus affecting antibiotic sorption to the clay surface. 305

    Furthermore, presence of ions in soil solution could impact antibiotic sorption also influencing 306

    antibiotic bioavailability and mobility in soils (Figueroa et al. 2004). However, since limited 307

    research has been conducted on the behavior of these chemicals in situ, results of pure clay 308

    studies provide insight into tetracycline behavior in soils. 309

    In addition to sorption to clay surfaces, tetracyclines intercalate the interlayer of 2:1 swelling 310

    clays (Kumar et al. 2005). For example, Chang et al. (2009a,b) show through X-ray diffraction 311

    (XRD) and Fourier transform-infrared (FTIR) analyses that tetracycline intercalates rectorite 312

    between pH 1.5 – 8.7. Similarly, OTC intercalates montmorillonite between pH 1.5 and 11.0, as 313

    evidenced by increased d-spacing observed through XRD analyses (Kulshrestha et al. 2004; 314

    Aristilde et al. 2013). In a study of OTC desorption in30soils with variable properties, the 315

    amounts desorbed in the Iredell Alfisol and the Sharkey Vertisol (37 and 53%, respectively) 316

    Page 14 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    15

    were noticeably less than the average 75 ± 14% desorbed from the larger set of soils (Jones et al. 317

    2005). The high proportion of shrink-swell clays in these soils, coupled with high Ca2+

    and Mg2+

    318

    contents, supports OTC intercalation, thereby yielding low recoveries (Jones et al. 2005; 319

    Aristilde et al. 2016). Aristilde et al. (2016) found that when Ca2+

    or Mg2+

    were present in 320

    solution, OTC adsorption to montmorillonite doubled. Furthermore, they detected a 120% 321

    increase in adsorbed OTC when introduced with Mg2+

    , accompanied by a 200% increase in OTC 322

    intercalation of montmorillonite layers. Intercalation of OTC into montmorillonite can be 323

    affected by ionic strength (Figueroa et al. 2004). Sorption edge experiments using three salt 324

    concentrations showed that while the OTC cationic species interacted more strongly with 325

    montmorillonite than the zwitterion, the cation was more sensitive to high ionic strength, as seen 326

    in a near 13-fold decrease in Kd when ionic strength increased from 10 mM to 510 mM 327

    (Figueroa et al. 2004). Although the adsorption capacity of kaolinite is far less than for swelling 328

    clays due to a lack of interlayer accessibility, the rate of adsorption is much higher since 329

    tetracycline is interacting only with the external surface and not intercalating (Li et al. 2010). 330

    Tetracycline adsorption to clay has been characterized by two kinetically different processes: a 331

    fast initial adsorption to outer surfaces of clays, followed by slower intercalation of swelling 332

    clays and into micropores (Sithole and Guy 1987a). 333

    The strong correlation between tetracycline sorption and effective cation exchange capacity 334

    (CEC) could result from electrostatic attraction between the tertiary amine group that is 335

    positively charged at environmentally relevant soil pHs and the negatively charged clay surface 336

    (Jones et al. 2005; Zhao et al. 2015). As this study also found a strong positive correlation 337

    between Kd and Mg2+

    , Ca2+

    , and Na+, cation bridging between negatively charged functional 338

    groups and base cations may contribute to the strong interaction between tetracyclines and 339

    Page 15 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    16

    smectite clays (Figueroa et al. 2004). Tetracyclines also bind to proteins and anionic silanol 340

    groups common on phyllosilicate surfaces (Sithole and Guy 1987a; Loke et al. 2002). Although 341

    kaolinite has a low CEC compared to 2:1 swelling clays, sorption of tetracyclines on kaolinite 342

    needs to be considered because it may contribute significantly to tetracycline sorption when 343

    present in high concentrations in soils. Furthermore, the pH-dependent surface charge on 344

    kaolinite influences tetracycline interactions with kaolinite, with maximum sorption on kaolinite 345

    at pH 7 when it is a zwitterion (Li et al. 2010). Although the net charge of the tetracycline 346

    zwitterion is zero, this species contributes significantly to total tetracycline sorption to soil clays 347

    because it is the dominant tetracycline species at the common soil pH values of between 348

    approximately 3.3 and 8. In addition to the dominant sorption mechanism of cation exchange 349

    which predominates on kaolinite surface sites (Zhao et al. 2015), surface complexation via H-350

    bonding is responsible for tetracycline sorption to kaolinite edge sites (Li et al. 2010; Zhao et al. 351

    2015). The size and physical conformation of tetracycline aligned with the kaolinite surface does 352

    not appear to influence sorption, but rather charge density shows the greatest effect (Li et al. 353

    2010). For example, Zhao et al. (2011) concluded that TC adsorption on kaolinite decreased with 354

    the increasing atomic radius and valence of metal cations in solution, suggesting that cations 355

    interfere with outer-sphere complexation between TC and kaolinite. 356

    As pH increases, tetracycline molecules become increasingly deprotonated, thus exhibiting a 357

    greater net negative charge and correspondingly weaker sorption to soils (Sassman and Lee 358

    2005; Peng et al. 2014; Zhao et al. 2015). This behavior is illustrated by decreasing Kd values 359

    with increasing pH (Figueroa et al. 2004; Sassman and Lee 2005). For example, the Kd of OTC 360

    sorption to Na-montmorillonite was approximately 7500 L eq-1

    at pH 5, declining to 361

    approximately 800 L eq-1

    by pH 8 (Figueroa et al. 2004). This decrease in sorption from pH 5 to 362

    Page 16 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    17

    8 reflects the decrease in zwitterion abundance in that pH range and the significance of the 363

    zwitterion species on total OTC sorption. Furthermore, total sorption of tetracyclines at any pH is 364

    a sum of the sorption of cationic species and zwitterionic species (Figueroa et al. 2004). 365

    Figureroa et al. (2004) concluded that although cationic OTC species comprise a small portion of 366

    total OTC at pH 5.5, the force of attraction between the cationic species and montmorillonite is 367

    20 times stronger than that of the zwitterion. Accordingly, 19% of total sorption was ascribed to 368

    the cation and 81% to the zwitterion. 369

    Tetracyclines can adsorb to soils via highly stable chelate complexes and ionic bridging with 370

    metals, including divalent cations Cu2+

    , Ca2+

    , and Mg2+

    (Loke et al. 2002; MacKay and 371

    Canterbury 2005; Sassman and Lee 2005; Avisar et al. 2010; Zhang et al. 2011). While 372

    tetracycline sorption decreases with increasing pH due to its deprotonation, sorption to alkaline 373

    soils is enhanced when those soils contain high Ca2+

    and Mg2+

    concentrations (Arias et al. 2007). 374

    Above pH 7 when anionic tetracycline species gain prominence, the presence of Ca2+

    in solution 375

    increases Kd values compared to tetracycline sorption in the absence of cations (Figueroa et al. 376

    2004). As mentioned in numerous studies, this enhanced tetracycline sorption at alkaline pH is 377

    attributed to cation bridging between the carboxylic residue of the tetracycline molecule and 378

    negatively charged clay surfaces (Avisar et al. 2010). However, the presence of Ca2+

    or Na+ in 379

    solution below pH 7 results in significantly less sorption of OTC to montmorillonite as a result of 380

    competition between positively charged OTC species and cations for sorption sites on the clay 381

    surface. In contrast, the presence of Cu2+

    at low soil pH (

  • Draft

    18

    though the latter may be the dominant mechanism (Gu and Karthikeyan 2008) (Fig. 3). Sorption 386

    of tetracyclines on SOM is highly pH-dependent between pH 2.5 and 10 due to variable charges 387

    on the three ionizable functional groups, with maximum tetracycline sorption at pH 4.3 when the 388

    zwitterionic form comprises 90% of total species. This species interacts primarily with 389

    carboxylic acid functional groups on SOM via cation exchange (Avisar et al. 2010). Below pH 390

    4.3, zwitterionic and cationic tetracycline species compete with H+

    for binding on humic acids 391

    and above 4.3, sorption decreases due to charge repulsion between the predominant anionic 392

    tetracycline species and deprotonated SOM (Gu et al. 2007). Tetracycline sorption to SOM is 393

    strongly influenced by the presence of multivalent cations (Fig. 3). For example, divalent cations 394

    such as Ca2+

    and Mg2+

    exert a significant effect on tetracycline sorption to SOM at pHs >5, 395

    leading to formation of ternary organic matter-M2+

    - tetracycline complexes (Gu et al. 2007). 396

    Bridging with multivalent cations has likewise been determined as a means of OTC sorption to 397

    manure in soil (Sithole and Guy 1987b; Loke et al. 2002). The Kd values of tetracyclines to 398

    manure thus cannot be determined by extrapolation from Kow or the fraction of organic carbon in 399

    soil, for Kd values are much higher than expected based on the negative log Kow values (Sithole 400

    and Guy 1987b; Loke et al. 2002; Jones et al. 2005). 401

    While sorption of tetracyclines on clay surfaces decreases with increasing pH, a greater 402

    overall decline in sorption was seen with increasing ionic strength (Sithole and Guy 1987a; 403

    Figueroa et al. 2004; Li et al. 2010). Tetracycline sorption is strongly influenced by the 404

    concentration of metals in solution, as evidenced by a more than 50% decrease in sorption to 405

    SOM when ionic strength increased 10-fold (Gu and Karthikeyan 2008). This correlation was 406

    especially strong for cationic tetracycline species, suggesting that cation exchange is the most 407

    significant sorption mechanism at low pH (Figueroa et al. 2004). However, surface complexation 408

    Page 18 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    19

    is a significant sorption mechanism for zwitterionic tetracycline species that predominate 409

    between pHs of approximately 3.3 and 8, indicated by their low sensitivity to ionic strength. 410

    Hysteretic behavior has been noted for humic acid-tetracycline complexes, whereby a large 411

    concentration of tetracycline bound irreversibly to humic acids via physical containment or 412

    changes in humic acid conformation (Gu et al. 2007). Hysteresis may decrease bioavailability, 413

    degradation, and mobility of tetracyclines in soils. 414

    Organic matter is not a primary influence on tetracycline sorption to soils, which must be 415

    taken into account when assessing the effect of land-application of manure (Teixido et al. 2012). 416

    This conclusion is illustrated in Chessa et al.’s (2016) study in which TC Kd values were nearly 417

    identical whether cow manure was applied to soil or not; this trend held for two distinct soils 418

    with moderate and high levels of manure application. Humic substances, either dissolved or 419

    bound to clay minerals, decrease tetracycline sorption, thus increasing its mobility (Avisar et al. 420

    2010). For example, a concentration of 10 mg/L humic acid reduced sorption of OTC to 421

    montmorillonite (Kulshrestha et al. 2004). Consequently, the potential exists for tetracyclines to 422

    migrate via surface and groundwaters in soils high in organic matter (Gu and Karthikeyan 2008). 423

    Jones et al. (2005) likewise saw a strong negative correlation between OTC sorption to 29 424

    different soils (≤ 4% organic carbon) and percent soil organic carbon. In contrast, strong OTC 425

    sorption to soils is sometimes attributed to high OM content. In a study of OTC sorption in three 426

    distinct soils, Peng et al. (2014) concluded that strong adsorption to soil B was due to its high 427

    OM content (42.7 g Kg-1

    ), in contrast to soil A and soil C with 10.4 and 13.4 g Kg-1

    , 428

    respectively. However, the low pH values (3.5 and 4.3, respectively) of soil B and C were 429

    perhaps greater determinants of OTC sorption, considering their comparable Kf values of 3590 430

    and 3640, respectively, in comparison to the low Kf (108) of soil A with a pH of 7.6. Similarly, 431

    Page 19 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    20

    Jones et al. (2005) suggest strong OTC sorption to the Burton soil may resulted in part from its 432

    high organic carbon content (8.93%) The low 3.2 pH of the Burton soil likely contributed to its 433

    strong OTC sorption. 434

    Interaction with Al, Fe, and Mn Oxides 435

    Interactions with metal oxides influence tetracycline behavior in soils. Carbonyl and 436

    tricarbonyl functional groups on tetracyclines complex strongly with Al and Fe atoms on edges 437

    of aluminum and iron hydrous oxides (HAO and HFO, respectively) (Gu and Karthikeyan 2005). 438

    However, tetracyclines compete for edge sites with SOM since the latter likewise strongly 439

    associates with HAO. Gu and Karthikeyan (2008) found that TC sorption onto HAO 440

    significantly declined with increasing humic acid content, with an increase from 0.81% to 1.52% 441

    organic carbon resulting in about a 40% decrease in TC sorption. Pils and Laird (2007) 442

    concluded that TC and CTC adsorb in decreasing order to clay minerals > humics > clay-humic 443

    complexes. Reduction in TC sorption in the presence of SOM may be due to competition for 444

    sorption sites, and also reversal in aluminum oxide surface charge resulting from SOM sorption, 445

    which would lead to repulsion between the oxide surface and deprotonated TC functional groups 446

    (Gu and Karthikeyan 2008). Tetracycline-SOM-HAO interactions were not obviously influenced 447

    by ionic strength, suggesting formation of inner sphere complexes (Gu and Karthikeyan 2008). 448

    Chen and Huang (2009, 2010, 2011) show that Mn2+

    , Cu2+

    , MnO2, and Al2O3 catalyze 449

    tetracycline transformations in soil including oxidations, dehydrations, isomerizations, and 450

    epimerizations. The pathways by which these changes occur vary according to the metal, 451

    resulting in various end products. A range of environmental fates for the tetracycline 452

    transformation products ensues. 453

    Page 20 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    21

    2.4. Tetracycline transport through soils 454

    Lysimeter-based studies are commonly used for examining leaching of tetracyclines through 455

    soil. In a study where fresh pig manure slurry amended with 0.87 kg ha-1

    OTC was added to 456

    lysimeter cores (60 cm long x 24 cm diameter) of a sandy loam soil, HPLC analyses did not 457

    detect OTC in any of the leachate samples (Blackwell et al. 2009). OTC was detected in the top 458

    5 cm of soil in two lysimeters at concentrations of 27 and 94 µg kg-1

    when manure was 459

    incorporated into the soil, and at concentrations of 19 and 41 µg kg-1

    in lysimeters where manure 460

    had been applied to the soil surface (Blackwell et al. 2009). Detection of OTC in soil, but not in 461

    leachate, is due to its strong sorption coefficient (though the authors did not provide a specific 462

    Kd) and persistence in soil, as similarly concluded in numerous other studies (Rabølle and Spliid 463

    2000; Hamscher et al. 2002, 2005; Aga et al. 2005; Kay et al. 2005). In a study involving 464

    repeated field applications of pig manure containing tetracycline (4 mg L-1

    ) and chlortetracycline 465

    (0.1 mg L-1

    ), TC and CTC were consistently detected in the top 30 cm of the soil, but there was 466

    no evidence of TC or CTC in the 30-90 cm depth, nor were any tetracyclines found in 467

    groundwater (detection limit was 50 ng L-1

    ) (Hamscher et al. 2002). 468

    Despite large attractive forces between tetracyclines and soils, they may be transported 469

    through the profile into groundwater via preferential flow processes (Kay et al. 2004; Aust et al. 470

    2008). Increasingly alkaline soil pH may also contribute to tetracycline leaching since the 471

    predominantly anionic species will not sorb as strongly to soils. In column studies investigating 472

    the effect of acid rain on TC mobility in soil following surface application of TC-spiked chicken 473

    manure (2000 µg TC kg soil-1

    ), TC was found in the top 20 cm at a concentration of 17.1 µg kg-1

    474

    after pH 3 rain, but was found at 0-40 cm depth at 6.57 µg kg-1

    after pH 7 rain, demonstrating its 475

    mobility in soil at higher pH (Pan and Chu 2017). In an extensive survey of three dairy farms in 476

    Page 21 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    22

    China, Zhou et al. (2013) analyzed the concentrations of 50 antibiotics in the fecal waste and 477

    wastewater. Tetracycline and CTC were detected in the fecal samples at concentrations of 16.7 478

    µg kg-1

    and 1450 µg kg-1

    , respectively. Tetracyclines (TC, OTC, and CTC) were consistently 479

    found in association with suspended particles in wastewater at all three dairy farms at 480

    concentrations ranging from 0 to 1710 µg kg-1

    , with CTC usually occurring in the highest 481

    concentrations (Zhou et al. 2013). Conversely, tetracyclines were rarely detected in the aqueous 482

    phase (Zhou et al. 2013). 483

    2.5. Influence of environmental conditions on tetracycline sorption and degradation 484

    Environmental conditions also influence tetracycline sorption and degradation in soils. 485

    Photolysis, heat, and microbial degradation have been implicated in tetracycline depletion in 486

    manure piles (Storteboom et al. 2007). Degradation of tetracyclines correlates to temperature, 487

    with greater persistence and accumulation in soils at low temperature (Hamscher et al. 2005; Li 488

    et al. 2010), and higher temperatures stimulating increased molecular reactivity and greater 489

    degradation (Loftin et al. 2008). Based on a high 86 ± 11% recovery of desorbed OTC from 490

    diverse soils, Jones et al. (2005) conclude that removal of OTC from solution is due to sorption 491

    mechanisms and not degradation. 492

    Concentration of OTC in manured dairy calf bedding declined rapidly in the initial 10 days, 493

    with a degradation half-life of < 10 d (De Liguoro et al. 2003). Moreover, OTC concentrations in 494

    30-d and 105-d dry-piled manure differed according to superficial, intermediate, or deep position 495

    in the pile. The superficial layer contained the highest concentrations, with amounts roughly 496

    twice those found in the deep layer. Concentrations in the intermediate layer were roughly four 497

    times less than the deep layer (De Liguoro et al. 2003). Potential factors responsible for the 498

    Page 22 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    23

    discrepancies between the positions in the manure pile include pH, antibiotic concentration, 499

    temperature, and moisture. The strong attraction of tetracyclines to soils at low temperatures 500

    implies that in those conditions, tetracyclines will be less available to degradation via abiotic or 501

    biological processes. The higher temperature in the intermediate layer likely facilitated 502

    tetracycline desorption and increased biological activity, thereby increasing its rate of 503

    degradation (De Liguoro et al. 2003). In a comparison of high- and low-intensity manure 504

    management, degradation rates of tetracyclines were greater in high-intensity management 505

    systems in which manure piles were watered regularly, turned weekly, and amended with equal 506

    parts dry leaves and fresh alfalfa (Storteboom et al. 2007). The increased degradation rate may 507

    result from a greater moisture level that would support an active microbial community, or from 508

    the increased oxygen level, supporting more efficient aerobic microbial degradation pathways. 509

    Following incorporation of piglet manure, which initially contained approximately 100-300 510

    mg CTC per kg dry wt manure, into the top 20 cm of two livestock field soils [a loamy sand soil 511

    (A) and a sandy soil (B)], CTC dissipation was assessed and its half-life was 25 d in soil A and 512

    34 d in soil B (Halling-Sørensen et al. 2005). Carlson and Mabury (2006) conducted a similar 513

    CTC dissipation study, but using dairy manure spiked with an initial CTC concentration of 514

    approximately 725 µg kg-1

    and incorporated into the upper 12 cm of a sandy loam soil. Despite 515

    the more than 100-fold less initial CTC concentration compared to the previous study, the 516

    average half-life was 24 d in the manure-amended plots and 21 d in the non-manure plots. 517

    Dissipation kinetics of OTC spiked in soils that were sterilized or non-sterilized, and under 518

    aerobic or anoxic conditions (Yang et al. 2009) yielded significant differences between 519

    treatments, and contrasting results from Carlson and Mabury (2006). For example, t1/2 of OTC in 520

    aerobic, non-sterile soil A (initial concentration of 10 mg kg-1

    ) was 37 d, compared to 99 d when 521

    Page 23 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    24

    the soil was sterilized, or 62 d was non-sterile soil A was under anoxic conditions. Increasing the 522

    initial OTC concentration in aerobic non-sterile soil A from 5 mg kg-1

    to 30 mg kg-1

    increased t1/2 523

    from 29 d to 56 d (Yang et al. 2009).The contrasting half-life figures found in different studies 524

    are due in part to compound structure, initial concentration, varying pH, soil redox status, 525

    temperature, and microbial community. , The half-life of CTC measured in spiked soil interstitial 526

    water in the dark at pH 3.0 and 4.1 was approximately 40 d, but dropped to about 2 d when pH 527

    was increased to 8.5 (Søeborg et al. 2004). This pH variability could explain the high dissipation 528

    rate of CTC (initial concentration 0.328 mg kg-1

    ) in spiked horse manure (pH 8.2-8.4), which had 529

    t1/2 of 5.1 to 8.4 days (Storteboom et al. 2007). The initial concentration of tetracyclines in beef 530

    feedlot manure of approximately 2 mg kg-1

    declined to below the 0.01 mg kg-1

    dry manure 531

    detection limit within 6 months, while initial tetracycline concentrations in dairy manure of 532

    approximately 0.4 mg kg-1

    declined to below the detection limit within 4 months (Storteboom et 533

    al. 2007). The discrepancy in initial tetracycline concentrations is expected since the beef cattle 534

    are routinely administered subtherapeutic levels of tetracyclines, while they are given to dairy 535

    cattle only for therapeutic purposes (Storteboom et al. 2007). 536

    2.6. Tetracycline degradation products 537

    Sparse information exists on the presence and fate of antibiotic degradation products in soils 538

    following manure application. In addition to the parent molecule, four epimers of tetracyclines 539

    can be excreted by livestock (Brambilla et al. 2007). Accordingly, Qiao et al. (2012) reported 540

    that epimers 4-epitetracycline (ETC), 4-epi-chortetracycline (ECTC), and 4-epioxytetracycline 541

    (EOTC) were the predominant degradation products detected in fresh and composted swine 542

    manure samples. Though present in fresh swine manure, none of the anhydrous tetracycline 543

    degradation products were detected in composted manure or treated soil samples, potentially 544

    Page 24 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    25

    because dehydrated products only form in very acidic conditions (Qiao et al. 2012). Similar 545

    results were achieved by Wu et al. (2011) who did not detect any of the 4-546

    epianhydrotetracyclines in composted swine manure. In leachate samples taken from manure-547

    amended soil columns, Kwon (2011) detected consistently elevated levels of CTC (0.075 ng mL-

    548

    1), almost continual release of ECTC, and in contrast to other reports, 4-epi-549

    anhydrochlotetracycline (EACTC), albeit at low and infrequently occurring levels. Kwon 550

    suggests detection of CTC and OTC parent molecules was due to increases in soil pH from 551

    manure application, leading to increased mobility of the compounds. The minor presence of 552

    epimers could be due to their formation requirement of acidic conditions. Considering that 553

    anhydrous products exhibit a greater biological toxicity than the parent molecules, longer 554

    composting times may mitigate potential biological risks from applying antibiotic-laden manure 555

    to soils (Halling-Sørensen et al. 2002; Qiao et al. 2012). In a study of California dairy farms, 556

    Watanabe et al. (2010) detected OTC, TC, and epi-tetracycline in surface soils following dairy 557

    manure application at concentrations of 25 µg kg-1

    , 8.8-105 µg kg-1

    , and 163 µg kg-1

    , 558

    respectively. 559

    Although CTC has a half-life of 5-84 days, half-lives of its degradation products can be as 560

    high as 400 d in soil interstitial water (Søeborg et al. 2004). These recalcitrant metabolites 561

    remain bioactive, thus able to continually exert selective pressure on soil microbes, which may 562

    be responsible for persistence of ARG in soils past the time when parent compounds have been 563

    depleted (Thiele-Bruhn 2003; Storteboom et al. 2007). Walczak and Xu (2011) likewise reported 564

    persistence of antibiotic resistance in dairy manure with time, raising concern for the increased 565

    distribution of ARG to the environment from manure application. 566

    2.7. Biological risk of tetracyclines 567

    Page 25 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    26

    Assessing the biological risk of tetracyclines in soils is not simple. Their strong sorption and 568

    persistence in soil would suggest low bioavailability (Tolls 2001), yet long residence times in 569

    soil have been linked to development of antibiotic resistance (Kümmerer 2009, 2010; Kyselková 570

    et al. 2015b). Allaire et al. (2006) saw no difference in CTC sorption between sterilized and non-571

    sterilized sandy loam or heavy clay soils, illustrating a possible lack of biological influence on 572

    tetracyclines during short residence times. In a study of plant uptake of various antibiotics from 573

    soils, Boxall et al. (2006) noted decreased growth of carrots and lettuce when grown in soils 574

    containing OTC residues. Modeling antibiotic transport in soils has been proposed as a method 575

    for assessing potential environmental risks of antibiotics, but Blackwell et al. (2009) found that 576

    predicted values produced from a commonly used pesticide fate model (FOCUS-PEARL v3.3.3) 577

    highly underestimated antibiotic transport detected in field lysimeters. A multitude of factors 578

    may contribute to the discrepancy, including interactions of manure colloids with antibiotics or 579

    fluctuating temperature and pH, underscoring the complexity of predicting antibiotic fate in the 580

    environment. Research suggests biological activity and toxicity of tetracyclines may be affected 581

    when they are chelated with metals (Loke et al. 2002); however, metal type, complex speciation, 582

    and organism tested all influence biological toxicity of chelated tetracyclines (Pulicharla et al. 583

    2017). 584

    In addition to the probable upsurge in antibiotic resistance of soil microbes, routine 585

    application of manure to agricultural soils has raised concerns about potential risks to soil 586

    microbial health, functioning, and community structure (Toth et al. 2011). For example, 587

    chlortetracycline has been found to negatively impact microbial iron reduction in soil, with an 588

    ED50 (effective dose with 50% inhibition of Fe-reduction) of 2.54 x 104 µg kg

    -1 (Thiele-Bruhn 589

    2005). However, when using lower concentrations of antibiotics, chlortetracycline demonstrated 590

    Page 26 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    27

    no influence on iron reduction (Toth et al. 2011). Other microbial ecosystem functions adversely 591

    influenced by antibiotics in soil include methanogenesis, nitrogen cycling, and sulfate reduction 592

    (Ding and He 2010). Microbial community diversity as determined by Shannon’s diversity 593

    index, as well as microbial substrate utilization, was significantly reduced with additions of OTC 594

    to soil, even at OTC concentrations as low as 1 µM (Kong et al. 2006). 595

    3. Conclusion 596

    Tetracycline antibiotics enter agricultural soils from land-application of dairy manure. 597

    Numerous factors affect their behavior in soils including soil pH, soil mineralogy, organic 598

    matter, and the microbial community. Complex interactions make it difficult to readily predict 599

    the fate of these CECs in soil. Soil solution pH, ionic strength, and soil organic matter content 600

    especially influence tetracycline fate in soils. Due to their amphoteric structures, the behavior of 601

    tetracyclines in soil is fundamentally governed by pH, which dictates the charge and physical 602

    properties of the compounds. In typical soil pHs, these compounds bind to soil, yet soil organic 603

    matter and metals greatly impact that association, and given the likelihood of high SOM and 604

    metal content in manure, these variables must be seriously considered. In general, tetracyclines 605

    are sorbed by soils at typical soil pHs. While this may be advantageous to waterways, its 606

    consequential promotion of antibiotic resistance in microbes poses a serious threat to human 607

    health. Moreover, tetracyclines degrade to form secondary products that exhibit yet additional 608

    properties distinct from their parent molecules. Given the variable interacting factors that affect 609

    tetracycline behavior in soil, additional research is needed to make well-informed management 610

    decisions regarding land-application of dairy manure. 611

    612

    Page 27 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    28

    4. Acknowledgements 613

    This work was supported by the Agriculture and Food Research Initiative competitive grant 614

    2013-67019-21375 from the USDA National Institute of Food and Agriculture. Financial support 615

    was also provided by the University of Idaho, Department of Plant, Soil, and Entomological 616

    Sciences. 617

    618

    Page 28 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    29

    References 619

    620

    Aga, D.S., O'Connor, S., Ensley, S., Payero, J.O., Snow, D., and Tarkalson, D. 2005. 621

    Determination of the persistence of tetracycline antibiotics and their degradates in manure-622

    amended soil using enzyme-linked immunosorbent assay and liquid chromatography− mass 623

    spectrometry. J. Agric. Food Chem. 53(18): 7165-7171. 624

    625

    Allaire, S.E., Del Castillo, J., and Juneau, V. 2006. Sorption kinetics of chlortetracycline and 626

    tylosin on sandy loam and heavy clay soils. J. Environ. Qual. 35: 969-972. 627

    628

    Arias, M., García-Falcón, M.B., García-Ríoc, L., Mejutod, J.C., Rial-Oterob, R., and Simal-629

    Gándarab, J. 2007. Binding constants of oxytetracycline to animal feed divalent cations. J. Food 630

    Eng. 78: 69–73. 631

    632

    Aristilde, L., Lanson, B., and Charlet, L. 2013. Interstratification patterns from the pH-dependent 633

    intercalation of a tetracycline antibiotic within montmorillonite layers. Langmuir 29(14): 4492-634

    4501. 635

    636

    Aristilde, L., Lanson, B., Miéhé-Brendlé, J., Marichal, C., and Charlet, L. 2016. Enhanced 637

    interlayer trapping of a tetracycline antibiotic within montmorillonite layers in the presence of 638

    Ca and Mg. J. Colloid Interface Sci. 464: 153-159. 639

    640

    Page 29 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    30

    Aust, M.O., Godlinski, F., Travis, G.R., Hao, X., McAllister, T.A., Leinweber, P., and Thiele-641

    Bruhn, S. 2008. Distribution of sulfamethazine, chlortetracycline and tylosin in manure and soil 642

    of Canadian feedlots after subtherapeutic use in cattle. Environ. Pollut. 156(3): 1243-1251. 643

    644

    Avisar, D., Primor, O., Gozlan, I., and Mamane, H. 2010. Sorption of sulfonamides and 645

    tetracyclines to montmorillonite clay. Water, Air, Soil Pollut. 209: 439-450. 646

    647

    Barkema, H.W., Schukken, Y.H., Lam, T.J.G.M., Beiboer, M.L., Benedictus, G., and Brand, A. 648

    1998. Management practices associated with low, medium, and high somatic cell counts in bulk 649

    milk. J. Dairy Sci. 81: 1917-1927. 650

    651

    Blackwell, P.A., Kay, P., Ashauer, R., and Boxall, A.B.A. 2009. Effects of agricultural 652

    conditions on the leaching behaviour of veterinary antibiotics in soils. Chemosphere 75: 13–19. 653

    654

    Boothe, D.M. 2015. Tetracyclines [online]. The Merck Veterinary Manual. Available from 655

    http://www.merckvetmanual.com/mvm/pharmacology/antibacterial_agents/tetracyclines.html 656

    [accessed 1 February 2014]. 657

    658

    Boxall, A.B.A., Kolpin, D.W., Halling-Sørensen, B., and Tolls, J. 2003. Are veterinary 659

    medicines causing environmental risks? Environ. Sci. Technol. 37: 286A-294A. 660

    661

    Boxall, A.B.A., Johnson, P., and Smith, E.J. 2006. Uptake of veterinary medicines from soils 662

    into plants. J. Agric. Food Chem. 54: 2288-2297. 663

    Page 30 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    31

    664

    Brambilla, G., Patrizii, M., De Filippis, S.P., Bonazzi, G., Mantovi, P., and Barchi, D. 2007. 665

    Oxytetracycline as environmental contaminant in arable lands. Anal. Chim. Acta 586: 326-329. 666

    667

    Carlson, J.C., and Mabury, S.A. 2006. Dissipation kinetics and mobility of chlortetracycline, 668

    tylosin, and monensin in an agricultural soil in Northumberland County, Ontario, Canada. 669

    Environ. Toxicol. Chem. 25: 1–10. 670

    671

    Carter, A.P., Clemons, W.M., Brodersen, D.E., Morgan-Warren, R.J., Wemberley, B.T., and 672

    Ramakrishnan, V. 2000. Functional insights from the structure of the 30S ribosomal subunit and 673

    its interactions with antibiotics. Nature 407: 340-348. 674

    675

    Chambers, L., Yang, Y., Littier, H., Ray, P., Zhang, T., Pruden, A., Strickland, M., and 676

    Knowlton, K. 2015. Metagenomic analysis of antibiotic resistance genes in dairy cow feces 677

    following therapeutic administration of third generation cephalosporin. PloS One 10(8): 678

    p.e0133764. 679

    680

    Chang, P.H., Jean, J.S., Jiang, W.T., and Li, Z. 2009a. Mechanism of tetracycline sorption on 681

    rectorite. Colloids Surf. A Physicochem. Eng. Asp. 339(1): 94-99. 682

    683

    Chang, P.H., Li, Z., Jiang, W.T. and Jean, J.S., 2009b. Adsorption and intercalation of 684

    tetracycline by swelling clay minerals. Appl. Clay Sci. 46(1): 27-36. 685

    686

    Page 31 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    32

    Chang, P.H., Jiang, W.T., Li, Z., Jean, J.S., and Kuo, C.Y. 2015. Antibiotic tetracycline in the 687

    environments- a review. Res. Rev.: J. Pharm. Anal. 4(3): 86-111. 688

    689

    Chee-Sanford, J.C., Mackie, R.I., Koike, S., Krapac, I.G., Lin, Y., Yannarell, A.C., Maxwell, S., 690

    and Aminov, R.I. 2009. Fate and transport of antibiotic residues and antibiotic resistance genes 691

    following land application of manure waste. J. Environ. Qual. 38: 1086-1108. 692

    693

    Chen, W.R., and Huang, C.H. 2009. Transformation of tetracyclines mediated by Mn (II) and Cu 694

    (II) ions in the presence of oxygen. Environ. Sci. Technol. 43: 401-407. 695

    696

    Chen, W.R., and Huang, C.H. 2010. Adsorption and transformation of tetracycline antibiotics 697

    with aluminum oxide. Chemosphere 79: 779-785. 698

    699

    Chen, W. R., and Huang, C. H. 2011. Transformation kinetics and pathways of tetracycline 700

    antibiotics with manganese oxide. Environ. Pollut. 159(5): 1092-1100. 701

    702

    Chessa, L., Pusino, A., Garau, G., Mangia, N.P., and Pinna, M.V. 2016. Soil microbial response 703

    to tetracycline in two different soils amended with cow manure. Environ. Sci. Pollut. Res. 23(6): 704

    5807-5817. 705

    706

    Chopra, I., and Roberts, M. 2001. Tetracycline antibiotics: mode of action, applications, 707

    molecular biology, and epidemiology of bacterial resistance. Microbiol Mol Biol Rev 65(2): 232-708

    260. 709

    710

    Page 32 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    33

    Daghrir, R., and Drogui, P. 2013. Tetracycline antibiotics in the environment: a review. Environ. 711

    Chem. Lett. 11: 209-227. 712

    713

    De Liguoro, M., Cibin, V., Capolongo, F., Halling-Sorensen, B., and Montesissa, C. 2003. Use 714

    of oxytetracycline and tylosin in intensive calf farming: evaluation of transfer to manure and soil. 715

    Chemosphere 52: 203-212. 716

    717

    Ding, C., and He, J. 2010. Effect of antibiotics in the environment on microbial populations. 718

    Appl. Microbiol. Biotechnol. 87(3): 925-941. 719

    720

    Feinman, S. E., and Matheson, J. C. 1978. Draft environmental impact statement: subtherapeutic 721

    antibacterial agents in animal feeds. Department of Health, Education, and Welfare, Public 722

    Health Service, Food and Drug Administration, Bureau of Veterinary Medicine. Rockville, MD. 723

    724

    Figueroa, R. A., Leonard, A., and Mackay, A.A. 2004. Modeling tetracycline antibiotic sorption 725

    to clays. Environ. Sci. Technol. 38: 476-483. 726

    727

    Granados-Chinchilla, F., and Rodríguez, C. 2017. Tetracyclines in food and feedingstuffs: from 728

    regulation to analytical methods, bacterial resistance, and environmental and health implications. 729

    J. Anal. Methods Chem. 2017. doi: 10.1155/2017/1315497 730

    731

    Gu, C., and Karthikeyan, K.G. 2005. Interaction of tetracycline with aluminum and iron hydrous 732

    oxides. Environ. Sci. Technol. 39: 2660-2667. 733

    Page 33 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    34

    734

    Gu, C., and Karthikeyan, K.G. 2008. Sorption of the antibiotic tetracycline to humic-mineral 735

    complexes. J. Environ. Qual. 37: 704-711. 736

    737

    Gu, C., Karthikeyan, K.G., Sibley, S.D., and Pedersen, J.A. 2007. Complexation of the antibiotic 738

    tetracycline with humic acid. Chemosphere 66: 1494-1501. 739

    740

    Gullberg, E., Cao, S., Berg, O.G., Ilbäck, C., Sandegren, L., Hughes, D., and Andersson, D.I. 741

    2011. Selection of resistant bacteria at very low antibiotic concentrations. PLoS Pathog. 7(7): 742

    p.e1002158. 743

    744

    Halling-Sørensen, B., Sengeløv, G., and Tjørnelund, J. 2002. Toxicity of tetracyclines and 745

    tetracycline degradation products to environmentally relevant bacteria, including selected 746

    tetracycline-resistant bacteria. Arch. Environ. Contam. Toxicol. 42: 263-271. 747

    748

    Halling-Sørensen, B., Jacobsen, A-M., Jensen, J., Sengelov, G., Vaclavik, E., and Ingerslev, F. 749

    2005. Dissipation and effects of chlortetracycline and tylosin in two agricultural soils: a field-750

    scale study in southern Denmark. Environ. Toxicol. Chem. 24: 802–810. 751

    752

    Hamscher, G., Pawelzick, H.T., Hoper, H., and Nau, H. 2005. Different behavior of tetracyclines 753

    and sulfonamides in sandy soils after repeated fertilization with liquid manure. Environ. Toxicol. 754

    Chem. 24: 861-868. 755

    756

    Page 34 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    35

    Hamscher, G., Sczesny, S., Hoper, H., and Nau, H. 2002. Determination of persistent 757

    tetracycline residues in soil fertilized with animal slurry by high-performance liquid 758

    chromatography with electrospray ionization tandem mass spectrometry. Anal. Chem. 74: 1509-759

    1518. 760

    761

    Hao, H., Cheng, G., Iqbal, Z., Ai, X., Hussain, H.I., Huang, L., Dai, M., Wang, Y., Liu, Z., and 762

    Yuan, Z. 2014. Benefits and risks of antimicrobial use in food-producing animals. Low-dose 763

    antibiotics: current status and outlook for the future. Front. Microbiol. 5, 288. doi: 764

    10.3389/fmicb.2014.00288. 765

    766

    Hou, J., Wan, W., Mao, D., Wang, C., Mu, Q., Qin, S., and Luo, Y. 2015. Occurrence and 767

    distribution of sulfonamides, tetracyclines, quinolones, macrolides, and nitrofurans in livestock 768

    manure and amended soils of Northern China. Environ. Sci. Pollut. Res. 22(6): 4545-4554. 769

    770

    Jones, A.D., Bruland, G.L., Agrawal, S.G., and Vasudevan, D. 2005. Factors influencing the 771

    sorption of oxytetracycline to soils. Environ. Toxicol. Chem. 24: 761-770. 772

    773

    Kay, P., Blackwell, P.A., and Boxall, A.B.A. 2004. Fate of veterinary antibiotics in macroporous 774

    tile drained clay soil. Environ. Toxicol. Chem. 23: 1136-1144. 775

    776

    Kay, P., Blackwell, P.A., and Boxall, A.B. 2005. A lysimeter experiment to investigate the 777

    leaching of veterinary antibiotics through a clay soil and comparison with field data. Environ. 778

    Pollut. 134(2): 333-341. 779

    Page 35 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    36

    780

    Kemper, N., Farber, H., Skutlarek, D., and Krieter, J. 2008. Analysis of antibiotic residues in 781

    liquid manure and leachate of dairy farms in Northern Germany. Agric. Water Manage. 95: 782

    1288-1292. 783

    784

    Kolpin, D.W., Furlong, E.T., Meyer, M.T., Zaugg, S.D., Barber, L.B., and Buxton, H.T. 2002. 785

    Pharmaceuticals, hormones, and other organic wastewater contaminants in U.S. streams, 1999-786

    2000: A national reconnaissance. Environ. Sci. Technol. 36: 1202-1211. 787

    788

    Kong, W.D., Zhu, Y.G., Fu, B.J., Marschner, P., and He, J.Z. 2006. The veterinary antibiotic 789

    oxytetracycline and Cu influence functional diversity of the soil microbial community. Environ. 790

    Pollut. 143(1): 129-137. 791

    792

    Kulshrestha, P., Giese, R.F., and Aga, D.S. 2004. Investigating the molecular interactions of 793

    oxytetracycline in clay and organic matter: insights on factors affecting its mobility in soil. 794

    Environ. Sci. Technol. 38: 4097-4105. 795

    796

    Kumar, K., Gupta, S.C., Chander, Y., and Singh, A.K. 2005. Antibiotic use in agriculture and its 797

    impact on the terrestrial environment. Adv. Agron. 87: 1-54. 798

    799

    Kümmerer, K. 2009. The presence of pharmaceuticals in the environment due to human use – 800

    present knowledge and future challenges. J. Environ. Manage. 90: 2354–2366. 801

    Page 36 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    37

    802

    Kümmerer, K. 2010. Pharmaceuticals in the environment. Annu. Rev. Environ. Resour. 35: 57–803

    75. 804

    805

    Kwon, J. 2011. Mobility of veterinary drugs in soil with application of manure compost. Bull. 806

    Environ. Contam. Toxicol. 87: 40-44. 807

    808

    Kyselková, M., Jirout, J., Vrchotová, N., Schmitt, H., and Elhottová, D. 2015a. Spread of 809

    tetracycline resistance genes at a conventional dairy farm. Front. Microbiol. 6(536). doi: 810

    10.3389/fmicb.2015.00536. 811

    812

    Kyselková, M., Kotrbová, L., Bhumibhamon, G., Chroňáková, A., Jirout, J., Vrchotová, N., 813

    Schmitt, H., and Elhottová, D. 2015b. Tetracycline resistance genes persist in soil amended with 814

    cattle feces independently from chlortetracycline selection pressure. Soil Biol. Biochem. 81: 259-815

    265. 816

    817

    Li, Z., Schulz, L., Ackley, C., and Fenske, N. 2010. Adsorption of tetracycline on kaolinite with 818

    pH-dependent surface charges. J. Colloid Interface Sci. 351: 254-260. 819

    820

    Liu, A., Fong, A., Becket, E., Yuan, J., Tamae, C., Medrano, L., Maiz, M., Wahba, C., Lee, C., 821

    Lee, K., and Tran, K.P. 2011. Selective advantage of resistant strains at trace levels of 822

    antibiotics: a simple and ultrasensitive color test for detection of antibiotics and genotoxic 823

    agents. Antimicrob. Agents Chemother. 55(3): 1204-1210. 824

    Page 37 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    38

    825

    Loftin, K.A., Adams, C.D., Meyer, M.T., and Surampalli, R. 2008. Effects of ionic strength, 826

    temperature, and pH on degradation of selected antibiotics. J. Environ. Qual. 37: 378-386. 827

    828

    Loke, M-L., Tjørnelund, J., and Halling-Sørensen, B. 2002. Determination of the distribution 829

    coefficient (log Kd) of oxytetracycline, tylosin A, olaquindox and metronidazole in manure. 830

    Chemosphere 48: 351-361. 831

    832

    MacKay, A.A., and Canterbury, B. 2005. Oxytetracycline sorption to organic matter by metal-833

    bridging. J. Environ. Qual. 34: 1964-1971. 834

    835

    Martinez-Carballo, E. Gonzalez-Barreiro, C., Scharf, S., and Gans, O. 2007. Environmental 836

    monitoring study of selected veterinary antibiotics in animal manure and soils in Austria. 837

    Environ. Pollut. 148: 570-579. 838

    839

    Pan, M., and Chu, L.M. 2017. Leaching behavior of veterinary antibiotics in animal manure-840

    applied soils. Sci. Total Environ. 579: 466-473. 841

    842

    Peng, F.J., Zhou, L.J., Ying, G.G., Liu, Y.S., and Zhao, J.L. 2014. Antibacterial activity of the 843

    soil‐bound antimicrobials oxytetracycline and ofloxacin. Environ. Toxicol. Chem. 33(4): 776-844

    783. 845

    846

    Page 38 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    39

    Peng, S., Zhou, B., Wang, Y., Lin, X., Wang, H., and Qiu, C. 2016. Bacteria play a more 847

    important role than nutrients in the accumulation of tetracycline resistance in manure-treated 848

    soil. Biol. Fertil. Soils 52(5): 655-663. 849

    850

    PHYSPROP. SRC, Inc. 2013. PHYSPROP Database [online]. Available from 851

    http://esc.syrres.com/fatepointer/search.asp [accessed 17 November 2016]. 852

    853

    Pils, J.R.V., and Laird, D.A. 2007. Sorption of tetracycline and chlortetracycline on K- and Ca- 854

    saturated soil clays, humic substances, and clay-humic complexes. Environ. Sci. Technol. 41: 855

    1928-1933. 856

    857

    Pol, M., and Ruegg, P.L. 2007. Treatment practices and quantification of antimicrobial drug 858

    usage in conventional and organic dairy farms in Wisconsin. J. Dairy Sci. 90: 249-261. 859

    860

    Pruden, A., Pei, R., Storteboom, H., and Carlson, K.H. 2006. Antibiotic resistance genes as 861

    emerging contaminants: studies in northern Colorado. Environ. Sci. Technol. 40: 7445-7450. 862

    863

    Pulicharla, R., Hegde, K., Brar, S.K., and Surampalli, R.Y. 2017. Tetracyclines metal 864

    complexation: Significance and fate of mutual existence in the environment. Environ. Pollut. 865

    221:1-14. 866

    867

    Qiao, M., Chen, W., Su, J., Zhang, B., and Zhang, C. 2012. Fate of tetracyclines in swine manure 868

    of three selected swine farms in China. J. Environ. Sci. 24: 1047-1052. 869

    Page 39 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    40

    870

    Rabølle, M., and Spliid, N.H. 2000 Sorption and mobility of metronidazole, olaquindox, 871

    oxytetracycline and tylosin in soil. Chemosphere 40: 715-722. 872

    873

    Salgado, M., Collins, M.T., Salazar, F., Kruze, J., Bolske, G., Soderlund, R., Juste, R., Sevilla, 874

    I.A., Biet, F., Troncoso, F., and Alfaro, M. 2011. Fate of Mycobacterium avium subsp. 875

    paratuberculosis after application of contaminated dairy cattle manure to agricultural soils. Appl. 876

    Environ. Microbiol. 77: 2122-2129. 877

    878

    Santamaría, J., López, L., and Soto, C.Y. 2011. Detection and diversity evaluation of tetracycline 879

    resistance genes in grassland-based production systems in Colombia, South America. Front. 880

    Microbiol. 2. 881

    882

    Sarmah, A.K., Meyer, M.T., and Boxall, A.B. 2006. A global perspective on the use, sales, 883

    exposure pathways, occurrence, fate and effects of veterinary antibiotics (VAs) in the 884

    environment. Chemosphere 65(5): 725-759. 885

    886

    Sassman, S.A., and Lee, L.S. 2005. Sorption of three tetracyclines by several soils: assessing the 887

    role of pH and cation exchange. Environ. Sci. Technol. 39: 7452-7459. 888

    889

    Page 40 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    41

    Sengeløv, G., Agersø, Y., Halling-Sørensen, B., Baloda, S.B., Andersen, J.S., and Jensen, L.B. 890

    2003. Bacterial antibiotic resistance levels in Danish farmland as a result of treatment with pig 891

    manure slurry. Environ. Int. 28: 587–595. 892

    893

    Sithole, B.B., and Guy, R.D. 1987a. Models for tetracycline in aquatic environments. I. 894

    Interaction with bentonite clay systems. Water, Air, Soil Pollut. 32: 303-314. 895

    896

    Sithole, B.B., and Guy, R.D. 1987b. Models for tetracycline in aquatic environments. II. 897

    Interactions with humic substances. Water, Air, Soil Pollut. 32: 315-321. 898

    899

    Søeborg, T., Ingerslev, F., and Halling-Sørensen, B. 2004. Chemical stability of chlortetracycline 900

    and chlortetracycline degradation products and epimers in soil interstitial water. Chemosphere 901

    57: 1515–1524. 902

    903

    Storteboom, H.N., Kim, S., Doesken, K.C., Carlson, K.H., Davis, J.G., and Pruden, A. 2007. 904

    Response of antibiotics and resistance genes to high-intensity and low-intensity manure 905

    management. J. Environ. Qual. 36: 1695-1703. 906

    907

    Sweetman, S. C. 2009. Martindale: The Complete Drug Reference. Pharmaceutical Press. 908

    London. 909

    910

    Page 41 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    42

    Teixido, M., Granados, M., Prat, M.D., and Beltran, J.L. 2012. Sorption of tetracyclines onto 911

    natural soils: data analysis and prediction. Environ. Sci. Pollut. Res. 19: 3087-3095. 912

    913

    Thiele-Bruhn, S. 2003. Pharmaceutical antibiotic compounds in soils- a review. J. Plant Nutr. 914

    Soil Sci. 166: 145-167. 915

    916

    Thiele-Bruhn, S. 2005. Microbial inhibition by pharmaceutical antibiotics in different soils—917

    dose-response relations determined with the iron(III) reduction test. Environ. Toxicol. Chem. 918

    24(4): 869–876. 919

    920

    Tolls, J. 2001. Sorption of Veterinary Pharmaceuticals in Soils: A Review. Environ. Sci. 921

    Technol. 35: 3397-3406. 922

    923

    Toth, J.D., Feng, Y., and Dou, Z. 2011. Veterinary antibiotics at environmentally relevant 924

    concentrations inhibit soil iron reduction and nitrification. Soil Biol. Biochem. 43: 2470-2472. 925

    926

    United States Department of Agriculture APHIS, VS, CEAH, National Animal Health 927

    Monitoring System. 2008a. Antibiotic use on U.S. dairy operations, 2002 and 2007 [online]. 928

    Available from 929

    https://www.aphis.usda.gov/animal_health/nahms/dairy/downloads/dairy07/Dairy07_is_Antibiot930

    icUse.pdf [accessed 28 January 2014]. 931

    932

    Page 42 of 53

    https://mc06.manuscriptcentral.com/er-pubs

    Environmental Reviews

  • Draft

    43

    United States Department of Agriculture APHIS, VS, CEAH, National Animal Health 933

    Monitoring System. 2008b. Dairy 2007 Part III: Reference of dairy cattle health and 934

    management practices in the United States, 2007 [online]. Available from 935

    https://www.aphis.usda.gov/animal_health/nahms/dairy/downloads/dairy07/Dairy07_dr_PartIII_936

    rev.pdf [accessed 6 May 2014]. 937

    938

    United States Department of Agriculture APHIS, VS, CEAH, National Animal Health 939

    Monitoring System. 2009. Dairy 2007 Part V: Changes in dairy cattle health and management 940

    practices in the United States, 1996- 2007 [online]. Available from 941

    https://www.aphis.usda.gov/animal_health/nahms/dairy/downloads/dairy07/Dairy07_dr_PartV_r942

    ev.pdf [accessed 6 May 2014]. 943

    944

    United States Environmental Protection Agency. 2013. Literature review of contaminants in 945

    livestock and poultry manure and implications for water quality [online]. EPA-820-R-13-002. 946

    Available from https://www.scribd.com/document/214717740/Literature-Review-of-947

    Contaminants-in-Livestock-and-Poultry-Manure-and-Implications-for-Water-Quality [accessed 948

    19 January 2014]. 949

    950

    United States Food and Drug Administration, Department of Health and Human Services. 2015. 951

    2014 Summary report on antimicrobials sold or distributed for use in food-producing animals 952

    [online]. Available from 953

    http://www.fda.gov/downloads/For