Toxicol. Sci. 2002 Dorman 18 25

Embed Size (px)

Citation preview

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    1/8

    Cytochrome Oxidase Inhibition Induced by Acute Hydrogen SulfideInhalation: Correlation with Tissue Sulfide Concentrations in the Rat

    Brain, Liver, Lung, and Nasal Epithelium

    David C. Dorman,1 Frederic J.-M. Moulin,2 Brian E. McManus, Kristen C. Mahle, R. Arden James, and Melanie F. Struve

    CIIT Centers for Health Research, 6 Davis Drive, PO Box 12137, Research Triangle Park, North Carolina 27709-2137

    Received June 28, 2001; accepted October 4, 2001

    Hydrogen sulfide (H2S) is an important brain, lung, and nose

    toxicant. Inhibition of cytochrome oxidase is the primary biochem-

    ical effect associated with lethal H 2S exposure. The objective of

    this study was to evaluate the relationship between the concentra-

    tion of sulfide and cytochrome oxidase activity in target tissues

    following acute exposure to sublethal concentrations of inhaledH2S. Hindbrain, lung, liver, and nasal (olfactory and respiratory

    epithelial) cytochrome oxidase activity and sulfide concentrations

    were determined in adult male CD rats immediately after a 3-h

    exposure to H2S (10, 30, 80, 200, and 400 ppm). We also deter-

    mined lung sulfide and sulfide metabolite concentrations at 0, 1.5,

    3, 3.25, 3.5, 4, 5, and 7 h after the start of a 3-h H 2S exposure to

    400 ppm. Lung sulfide concentrations increased during H2S expo-

    sure and rapidly returned to endogenous levels within 15 min after

    the cessation of the 400-ppm exposure. Lung sulfide metabolite

    concentrations were transiently increased immediately after the

    end of the 3-h H2S exposure. Decreased cytochrome oxidase ac-

    tivity was observed in the olfactory epithelium following exposure

    to > 30 ppm H 2S. Increased olfactory epithelial sulfide concen-

    trations were observed following exposure to 400 ppm H2S. Hind-brain and nasal respiratory epithelial sulfide concentrations were

    unaffected by acute H2S exposure. Nasal respiratory epithelial

    cytochrome oxidase activity was reduced following acute exposure

    to > 30 ppm H2S. Liver sulfide concentrations were increased

    following exposure to > 200 ppm H2S and cytochrome oxidase

    activity was increased following inhalation exposure to > 10 ppm

    H2S. Our results suggest that cytochrome oxidase inhibition is a

    sensitive biomarker of H2S exposure in target tissues, and sulfide

    concentrations are unlikely to increase postexposure in the brain,

    lung, or nose following a single 3-h exposure to < 30 ppm H2S.

    Key Words: hydrogen sulfide; pharmacokinetics; cytochrome

    oxidase; nasal toxicity; rat; inhalation.

    Hydrogen sulfide (H2S) is a colorless gas with a character-

    istic rotten-egg odor. In nature, H 2S is produced primarily by

    the decomposition of organic matter and is found in natural

    gas, petroleum, volcanic, and sulfur-spring emissions. Hydro-

    gen sulfide is associated with more than 70 types of industries,

    including artificial fiber synthesis, food production, paper and

    pulp manufacture, roofing, sewage treatment, and swine con-

    tainment (Donham et al., 1982; Hall and Rumack, 1997; Hoi-

    dalet al., 1986; Jaakkolaet al., 1990; Wattet al., 1997). Lethal

    human exposure to H2S is usually associated with individuals

    working within heavily contaminated confined spaces (e.g.,

    sewers, manure pits) and with the oil and sour gas industries

    (Arnoldet al., 1985; Guidotti, 1994; Kilburn, 1993). The toxic

    effects of sublethal doses have been much less characterized. A

    recent review of the adverse health effects from H 2S exposure

    is available (ATSDR, 1999).

    The primary mechanism for the toxic action of H 2S is direct

    inhibition of cytochrome oxidase, an enzyme critical for mito-

    chondrial respiration (Khan et al., 1990; Nicholls and Kim,

    1982). Tissues with high oxygen demand (e.g., brain and heart)

    are especially sensitive to disruption of oxidative metabolism

    by H2S (Ammann, 1986). Human exposure to H2S results inconcentration-dependent toxicity in the respiratory, cardiovas-

    cular, and nervous systems. Acute human exposure to rela-

    tively low concentrations ( 50 ppm) of H2S results in ocular

    and respiratory mucous membrane irritation leading to nasal

    congestion, pulmonary edema, and a syndrome known as gas

    eye, which is characterized by corneal inflammation (ATSDR,

    1999; Reiffenstein et al., 1992). Despite the strong character-

    istic odor associated with H 2S, many exposed individuals are

    unaware of its presence because their sense of smell is severely

    impaired following exposure to 150 ppm H2S. Acute human

    exposure to high concentrations of H2S (e.g., 500 ppm)

    results in a rapid onset of respiratory paralysis and uncon-

    sciousness that can result in death within minutes (Beauchampet al., 1984). Persistent sequelae of H2S poisoning are often

    related to the olfactory system and may include hyposmia,

    dysosmia, and phantosmia (Hirsch and Zavala, 1999; Kilburn,

    1997).

    Animal studies confirm that the olfactory system is espe-

    cially sensitive to H2S inhalation. Acute exposure of rats to

    moderately high concentrations of H2S ( 80 ppm) resulted in

    regeneration of the nasal respiratory mucosa and full thickness

    1 To whom correspondence should be addressed. Fax: (919) 558-1300.

    E-mail: [email protected] Present address: Bristol-Myers Squibb Pharmaceutical Research Institute,

    PO Box 5400, Princeton, NJ 08543-5400.

    TOXICOLOGICAL SCIENCES65, 1825 (2002)

    Copyright 2002 by the Society of Toxicology

    18

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    2/8

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    3/8

    H2S exposure. Methods used to generate and characterize the H2S expo-

    sure atmospheres have been previously described (Struve et al.,2001). Briefly,

    gas cylinders containing 5% (50,000 ppm) H 2S in nitrogen were purchased

    from Holox Gases (Cary, NC). Nose-only exposures were conducted using rat

    nose-only tubes and a nose-only system with 52 exposure ports (Cannon et al.,

    1983). Total air flow in the nose-only units was adjusted to provide approxi-

    mately 0.5 l/min per animal port. Hydrogen sulfide was metered through mass

    flow controllers (MKS Instruments, Andover, MA) and mixed with the nose-

    only unit air supply to provide the desired target H 2S concentration. ExposureH2S concentrations were determined by gas chromatography-FPD (Hewlett-

    Packard model 6890 with a GS-Q 30 meter 0.53 m Alltech column) at least

    six times during each 3-h exposure. The generation system was operated by the

    Andover Infinity control system (Andover Controls Corporation, Andover,

    MA). Animals were exposed to 0, 10, 30, 80, 200, or 400 ppm H 2S for 3 h.

    Cytochrome oxidase activity. Cytochrome oxidase activity was evaluated

    by determining the rate of oxidation of reduced ferricytochrome c using

    methods described by Weyant et al.(1988). Bovine-derived ferricytochrome c

    was initially dissolved (10 mg/ml) in a 0.01 M sodium phosphate buffer (pH

    7.0) and then reduced by the addition of ascorbic acid (2.4 mg/ml) for 24 h.

    Excess ascorbate was removed by equilibrium dialysis in a 0.01 M sodium

    phosphate buffer (pH 7.0) using 3,500 molecular weight cutoff tubing

    (Spectrum Medical Industries, Los Angeles, CA). Three changes of buffer

    were performed over a 24-h period. The assay reagent contained 0.7 ml (7 mg)

    reduced cytochrome c, 1 ml sodium phosphate buffer (pH 7.0), and 8.3 mldistilled water. The degree of reduction of the final assay mix was measured

    using a Beckman DV 650 UV/VIS spectrophotometer (Fullerton, CA), and the

    assay reagent was considered fully reduced if the A 550/A565 ratio was greater

    than 6.5, as specified by the product insert from Worthington Biochemical

    Corporation (Lakewood, NJ).

    The following modifications were made to the methods described by Wey-

    ant et al. (1988) to accommodate the use of a COBAS FARA II analyzer

    (Roche Diagnostic System, Somerville, NJ). Representative 50- to 200-mg

    tissue samples were diluted 10-fold with a 0.25 M sucrose buffer and homog-

    enized using an ultrasonic sonifier. Tissue homogenates were centrifuged

    (3000 g for 10 min at 4C). The supernatant was removed and added to an

    equivalent amount of 0.25 M sucrose buffer. The sample was then recentri-

    fuged (3000 g for 10 min) and the resulting supernatant used for the

    cytochrome oxidase assay. Enzyme activity was measured by monitoring the

    oxidation of reduced cytochrome c at 550 nm. Absorbance readings were taken

    following a 10-s incubation time and at 5-s intervals for 90 s. Total protein

    concentration within each sample was analyzed with the COBAS FARA II

    spectrophotometer using commercially available reagents (Roche Diagnostic

    System, Somerville, NJ).

    Determination of tissue sulfide concentrations. Hindbrain, liver, lung,

    and nasal epithelium samples from control rats and animals exposed to H 2S

    were evaluated for sulfide content. Tissue samples (50150 mg) were sec-

    tioned directly from frozen tissues, weighed, and placed into a clear glass

    crimp-top vial with molded conical bottom (Sun International, Wilmington,

    NC). The vials were sealed using a teflon septum, and 1 l tetraethylammo-

    nium hydroxide (TEAH, 35% aqueous solution, SACHEM, Austin, TX) per

    milligram of sample was added to the vial to digest the sample. The TEAH was

    added to the vial with a gastight syringe (Hamilton, Reno, NV) in order to

    minimize loss of H2S due to volatilization. Samples were centrifuged for 5 min

    at 3000

    g(4C) and were then kept at room temperature for 24 h to completethe sample digestion. After 24 h, 8 l of 28 mM NaOH per milligram sample

    were injected into the vial, and the samples were centrifuged again at 3000

    g for 5 min. A 100-l sample of the supernatant was then added to 400 l of

    28 mM NaOH in a polypropylene ConSert vial (Sun International, Wilming-

    ton, NC) to complete a 50-fold dilution of the tissue sample. The diluted

    supernatant sample was then injected into the liquid chromatography system.

    Sulfide and its metabolites were separated by high-performance liquid

    chromatography (HPLC) using methods adapted from Mitchell et al. (1993)

    and Rocklin and Johnson (1983). The liquid chromatogram consisted of a

    Model 580 dual-piston solvent delivery module (ESA Inc., Chelmsford, MA),

    a pulse-dampener, a Waters 717 plus refrigerated autosampler (Millipore

    Corporation, Milford, MA), and an IONPAC AS15 analytical column (4

    250 mm, Dionex Corporation, Sunnyvale, CA) with an IONPACAG15 guard

    column. Sulfide was detected using a Coulochem II electrochemical detector

    (ESA Inc., Chelmsford, MA) equipped with a model 5020 guard cell and a

    model 5040 amperometric analytical cell with silver target. The applied po-

    tentials of the guard cell and analytical cell were 584 and 50 mV, respectively.

    The output range of the ESA 5040 analytical cell was set at 20 nA/V.

    Lung sulfide metabolites were detected by a Dionex CD20 conductivitydetector (Dionex Corporation, Sunnyvale, CA). The conductivity detector was

    equipped with an anion self-regenerating suppressor (ASRS-Ultra 4 mm,

    Dionex Corporation) in recycle mode. The range of the conductivity detector

    was 3 s, and the anion self-regenerating suppressor was set at 300 mA. The

    data were acquired and integrated by a Baseline 810 chromatography work-

    station (Waters, Millipore Corporation, Milford, MA).

    Analytical-grade reagents were used, and all standards and eluents were

    prepared using distilled, deionized water with a specific resistance of 17.8

    megohm-cm. Sulfide, sulfite, sulfate, and thiosulfate were eluted isocratically

    at a flow rate of 1.5 ml/min using helium-degassed 28 mM NaOH as the mobile

    phase. Under these conditions, elution times for sulfide, sulfite, sulfate, and

    thiosulfate were 4, 8, 10, and 35 min, respectively. Tissue concentrations were

    determined from the linear regression (r2 0.95) of a calibration curve based

    on aqueous samples within a range of 550 ppb for sulfide and 110 ppm for

    sulfite, sulfate, and thiosulfate. The assay detection limit for sulfide was 1

    ng/ml, corresponding to 0.05 g/g tissue. Recovery rates for sulfide, sulfite,

    sulfate, and thiosulfate were 88 5%, 85 5%, 71 10%, and 102 8%,

    respectively, based on recovery rates obtained from control rat liver samples

    (n 35) spiked with known amounts of each analyte of interest.

    Statistics. Unless otherwise noted, data are reported as means SEM. All

    statistical analyses were performed using a standard statistical package (JMP,

    SAS Institute Inc., Cary, NC). Time-course and dose-response data were

    evaluated by one-way analysis of variance (ANOVA) followed by a compar-

    ison with the preexposure (control) group using Dunnetts test. The distribution

    of all data was tested for normality using the Shapiro-Wilk test before analysis.

    Linear regression correlations were performed according to standard statistical

    procedures and tested by analysis of variance. For all tests, a p value of 0.05

    or less was considered significant.

    RESULTS

    Tissue Sulfide Concentrations following H2S Exposure

    Sulfide was observed in all control (preexposure) tissue

    samples. Mean lung and liver sulfide concentrations in control

    animals were 0.54 0.03 and 0.55 0.11 g/g, respectively.

    As expected, lung sulfide concentrations increased during ex-

    posure to 400 ppm H 2S, reaching peak concentrations at the

    end of the 3-h H 2S exposure (Fig. 2). Lung sulfide concentra-

    tions rapidly decreased to preexposure levels within 15 min

    after the end of the H2S exposure (Fig. 2). Significantly in-

    creased end-of-exposure lung sulfide concentrations were ob-

    served following exposure to 80 ppm H 2S (Table 1). Asignificant positive nonlinear relationship between the H2S

    exposure concentration and the end-of-exposure lung sulfide

    concentration was observed. Significantly increased end-of-

    exposure liver sulfide concentrations were also observed fol-

    lowing exposure to 200 ppm H 2S (Table 1). End-of-expo-

    sure hindbrain sulfide concentrations were unaffected by H 2S

    exposure (Table 2). Although end-of-exposure nasal respira-

    tory epithelium sulfide concentrations were elevated following

    20 DORMAN ET AL.

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    4/8

    exposure to 400 ppm H 2S (Table 2), the observed increase wasnot statistically significant (p 0.065). In contrast, end-of-

    exposure olfactory epithelium sulfide concentrations were sig-

    nificantly increased following exposure to 400 ppm H 2S (Table

    2). Neither lung nor hindbrain sulfide concentration was in-

    creased following subchronic exposure to 80 ppm H2S (Ta-

    ble 3).

    Lung Sulfite, Sulfate, and Thiosulfate Concentrations

    Mean preexposure lung sulfite, sulfate, and thiosulfate con-

    centrations were 359 18, 140 7, and 505 27 g/g,respectively. Significantly increased lung sulfite, sulfate, and

    thiosulfate concentrations were observed in rats exposed to 400

    ppm H2S and occurred 15 min after the end of the 3-h H 2S

    exposure (Fig. 3). Lung sulfite, sulfate, and thiosulfate concen-

    FIG. 2. Mean ( SEM) lung sulfide concentrations before, during, and

    after a 3-h inhalation exposure to 400 ppm H 2S (n 6 rats/time point).

    *Indicates statistically different from preexposure control values (p 0.05).

    TABLE 1

    Mean ( SEM) Lung and Liver Sulfide Concentrations and

    Cytochrome Oxidase Activity Immediately After the End of a

    3-Hour H2S Exposure

    H2S exposure Sulfide concentration Cytochrome oxida se

    Lung

    0 0.54 0.03 1.76 0.02

    10 0.43 0.03 1.65 0.06

    30 0.64 0.05 1.56 0.06*

    80 0.78 0.04* 1.46 0.04*

    200 0.97 0.02* 1.43 0.06*

    400 0.88 0.10* 1.16 0.04*

    Liver0 0.55 0.11 2.06 0.10

    10 1.11 0.22 2.74 0.10*

    30 0.74 0.08 3.04 0.11*

    80 0.86 0.06 2.79 0.23*

    200 1.55 0.20* 2.85 0.19*

    400 1.55 0.42* 2.87 0.12*

    Note. H2S exposure is given in ppm, sulfide concentration in g/g, cyto-

    chrome oxidase in U/mg protein.

    *Indicates statistically different from control values (p 0.05).

    TABLE 2

    Mean ( SEM) Hindbrain and Nasal Epithelium Sulfide Con-

    centrations Immediately After the End of a Single 3-Hr H2S

    Exposure and Cytochrome Oxidase Activity Following One or

    Five 3-Hour H2S Exposures

    H2S exposure

    Single exposure

    Five

    exposures

    Sulfide

    concentration

    Cytochrome

    oxidase

    Cytochrome

    oxidase

    Hindbrain

    0 1.21 0.05 2.14 0.12 ND

    200 1.12 0.05 2.55 0.26 ND

    400 1.14 0.04 2.28 0.31 ND

    Respiratory epithelium

    0 1.73 0.14 1.23 0.05 1.02 0.17

    30 ND 0.60 0.08* 0.83 0.11

    80 ND 0.50 0.08* 0.60 0.19

    200 1.37 0.11 0.92 0.02* 0.74 0.07

    400 2.73 0.77 0.94 0.02* 0.86 0.19

    Olfactory epithelium

    0 1.42 0.11 1.20 0.06 1.13 0.0930 ND 1.01 0.07* 0.75 0.13*

    80 ND 0.99 0.07* 0.84 0.11*

    200 1.25 0.06 0.92 0.03* 0.70 0.12*

    400 2.07 0.33* 0.92 0.04* 0.51 0.01*

    Note. H2S exposure is given in ppm, sulfide concentration in g/g, cyto-

    chrome oxidase in U/mg protein. ND, not determined.

    *Indicates statistically different from control values (p 0.05).

    TABLE 3

    Mean ( SEM) Hindbrain and Lung Sulfide Concentrations

    Immediately After the End of a Subchronic (70-day) H 2S Expo-

    sure

    H2S exposure (ppm)

    Sulfide

    concentration (g/g)

    Cytochrome oxidase

    (U/mg protein)

    Hindbrain

    0 2.89 0.21 2.48 0.23

    10 2.80 0.26 2.21 0.14

    30 2.72

    0.30 2.17

    0.0480 2.98 0.22 2.12 0.02

    Lung

    0 1.01 0.14 1.04 0.04

    10 0.98 0.12 1.05 0.03

    30 0.96 0.12 0.95 0.02

    80 0.90 0.15 0.87 0.02*

    Note. H2S exposure is given in ppm, sulfide concentration in g/g, cyto-

    chrome oxidase in U/mg protein.

    *Significantly different from control value (p 0.05).

    21HYDROGEN SULFIDE TOXICOKINETICS AND TOXICODYNAMICS

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    5/8

    trations rapidly decreased to preexposure levels within minutesafter this transient increase (Fig. 3).

    Cytochrome Oxidase Activity

    Decreased lung cytochrome oxidase activity was observed

    following exposure to 30 ppm H2S (Table 1). End-of-

    exposure lung cytochrome oxidase activity was inversely lin-

    early correlated (r2 0.65) with lung sulfide concentration

    (p 0.0001, ANOVA). Liver cytochrome oxidase activity was

    significantly increased to approximately 130168% of preex-

    posure levels in all H 2S-exposed animals (Table 1). Hindbrain

    cytochrome oxidase activity was unaffected by H 2S inhalation

    (Table 2). Decreased olfactory and nasal respiratory epitheliumcytochrome oxidase activities were observed following a single

    3-h exposure to 30 ppm H 2S (Table 2). Repeated (5-day)

    exposure to H2S also resulted in significant cytochrome oxi-

    dase inhibition in the olfactory epithelium (Table 2). Repeated

    (5-day) exposure to H2S did not affect cytochrome oxidase

    activity in the nasal respiratory nasal epithelium (Table 2).

    Subchronic exposure to 80 ppm H 2S resulted in reduced cyto-

    chrome oxidase activity in the lung but not the hindbrain

    (Table 3).

    DISCUSSION

    This study examined the toxicokinetics of H2S in rats fol-lowing acute exposure to sublethal concentrations of the gas.

    The highest H2S exposure concentration used in our study (400

    ppm) occurs with acute accidental human poisonings and is

    associated with olfactory paralysis in humans and olfactory

    epithelial necrosis in laboratory animals (ATSDR, 1999; Bren-

    nemanet al.,2001; Lopez et al.,1988b). The lowest exposure

    concentration used in our study (10 ppm) equals the current

    TLV-TWA recommended by the American Conference of

    Governmental Industrial Hygienists. In contrast, ambient at-

    mospheric H2S concentrations range from 0.01 to 50 ppb,

    depending on the proximity of the sampling site to tidal flats,

    marshes, anaerobic soils, and other environmental sources of

    H2S (Warneck, 1988; Graedel et al., 1986).

    Hydrogen sulfide is normally present in mammalian tissues,

    and some evidence suggests that it is required for certain types

    of nerve transmission (Kimura, 2000). Literature values forendogenous levels of sulfide are variable and depend on the

    procedures used to extract the sulfide from the tissue and the

    analytical chemical methods used to quantify this metabolite

    (Goodwinet al.,1989; Kageet al.,1988; Mitchellet al.,1993).

    Special care must be taken to minimize the loss of free sulfide

    from the tissue sample due to the volatility of this gas. We used

    a strong base (TEAH) not only to digest our tissue samples, but

    also to minimize evaporative losses of H 2S. Another advantage

    of our analytical method is that the use of an amperometric

    analytical cell and a conductivity detector allowed us to simul-

    taneously detect and quantify sulfide and sulfide metabolites in

    the same sample of tissue. Endogenous tissue sulfide concen-

    trations determined with our analytical methods are similar to

    those reported in the literature. For example, brain sulfide

    concentrations observed in naive rats (1.21 0.05 g/g) from

    our acute study are comparable to values (1.94 0.24 g/g)

    reported by Warenycia and coworkers (1990), although the

    results from our 70-day exposure are somewhat higher. Back-

    ground lung tissue sulfate concentrations observed in our study

    (140 6.5 g/g) were approximately 2.5-fold higher than

    those observed by Rozman and coworkers (1992).

    Our interest in the lung was stimulated by possible portal-

    of-entry effects associated with H 2S inhalation and the known

    relationship between H2S inhalation and pulmonary edema and

    fibrinocellular alveolitis (Lopezet al.,1988a). As expected, weobserved that end-of-exposure lung sulfide concentrations were

    highly correlated with the amount of H2S in the exposure

    atmosphere, and elevated lung sulfide concentrations were

    observed following a single 3-h H2S exposure to 80 ppm.

    End-of-exposure lung sulfide concentrations were not in-

    creased in rats exposed subchronically to 80 ppm. In con-

    trast, decreased cytochrome oxidase activity was observed in

    the lung after a 3-h exposure to 30 ppm H2S as well as

    following subchronic exposure to 80 ppm H 2S. This finding

    indicates that cytochrome oxidase inhibition is a more sensitive

    biomarker of H2S exposure than tissue sulfide concentrations.

    Lung sulfide concentrations rapidly returned to preexposure

    levels within minutes after the end of a 3-h exposure to 400ppm H2S, suggesting that rapid pulmonary elimination or me-

    tabolism of sulfide occurs. An accumulation of sulfide metab-

    olites was not observed in the lung during the 3-h H 2S expo-

    sure. Transient increases in lung sulfite, sulfate, and thiosulfate

    concentrations were observed, however, immediately after the

    end of the 400-ppm H2S exposure. This increase in sulfide

    metabolite concentrations occurred coincidentally with the

    rapid decrease in lung sulfide concentration. This observation

    FIG. 3. Mean ( SEM) lung sulfite, thiosulfate, and sulfate concentrations

    before, during, and after a 3-h inhalation exposure to 400 ppm H2S (n 6

    rats/time point). *Indicates statistically increased lung sulfite, thiosulfate, and

    sulfate concentrations when compared with preexposure (control) values (p

    0.05).

    22 DORMAN ET AL.

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    6/8

    suggests that the detoxification of sulfide to sulfate may be-

    comes less effective as the concentration of sulfide increases in

    blood and other tissues due to H2S exposure (Fischer et al.,

    2000). A similar pharmacokinetic pattern has been observed in

    mice exposed to benzene, where competitive inhibition of an

    intermediate metabolite (phenol) occurs and formation of an-

    other metabolite (hydroquinone) is delayed until after the ben-

    zene inhalation ends (Medinsky et al., 1996; Rickert et al.,1979). Additional studies will be required to confirm our

    hypothesis that competitive inhibition of sulfide metabolism

    occurs during H2S inhalation.

    We also observed increased sulfide concentrations and cy-

    tochrome oxidase inhibition in the upper respiratory tract from

    H2S-exposed rats. Although our data showed that olfactory, but

    not respiratory, epithelial sulfide concentrations were signifi-

    cantly elevated following exposure to 400 ppm H2S, the ob-

    served differences in end-of-exposure tissue sulfide concentra-

    tions are unlikely to be toxicologically significant. For

    example, end-of-exposure olfactory epithelium sulfide concen-

    trations following a 3-h exposure to 400 ppm H 2S were 146%

    of levels observed in unexposed animals, whereas end-of-

    exposure respiratory epithelium sulfide concentrations ob-

    served in the same animals were 158% of control levels.

    Cytochrome oxidase activity was significantly decreased

    within the olfactory and respiratory nasal epithelium immedi-

    ately after a single 3-h exposure to 30 ppm H2S. Olfactory

    cytochrome oxidase activity was also significantly decreased to

    4566% of control levels following 5 consecutive days of

    exposure to 30 ppm H2S. Repeated (5-day) H2S exposure

    did not change cytochrome oxidase activity in the respiratory

    nasal epithelium. This observation is consistent with our recent

    studies that showed that regeneration of the nasal respiratory

    mucosa occurs rapidly during this 5-day period, whereas ne-crosis of the olfactory mucosa increased in severity (Brenne-

    man et al., 2001). Our data suggest that the regenerated respi-

    ratory epithelium becomes resistant to H2S-induced

    cytochrome oxidase inhibition. It should be noted that our

    dose-response data for cytochrome oxidase activity in the nasal

    tissue demonstrates some inconsistencies. For example, H2S-

    induced inhibition of nasal respiratory epithelial cytochrome

    oxidase activity was greater in rats exposed to either 30 or 80

    ppm than in rats acutely exposed to 200 ppm H2S. This

    unusual dose-response relationship may indicate that some

    compensatory changes are occurring in the epithelium in re-

    sponse to H2S exposure. Furthermore, regional differences in

    sulfide delivery within the olfactory epithelium are correlatedwith the development of nasal pathology at this site (Moulin et

    al., 2001). Our sampling procedure does not allow us to detect

    regional differences in H2S delivery to the olfactory epithe-

    lium, as we pooled the entire olfactory epithelium into one

    sample.

    It is reasonable to question whether cytochrome oxidase

    inhibition is a mode of action for H 2S-induced olfactory pa-

    thology. Our laboratory has shown that acute inhalation expo-

    sure of male rats to 400 ppm H 2S results in severe mitochon-

    drial swelling in degenerating olfactory neurons within the

    olfactory epithelium (Brenneman et al.,2001). This ultrastruc-

    tural lesion is consistent with, but not specific for, H2S-induced

    anoxic cell injury due to cytochrome oxidase inhibition. These

    data provide strong evidence that cytochrome oxidase inhibi-

    tion may indeed play a critical role in H2S-induced olfactory

    pathology. When considered together, our data suggest that theolfactory neuroepithelium is intrinsically more sensitive than

    the nasal respiratory epithelium to H2S-induced cytochrome

    oxidase inhibition. This result is not unexpected, as neurons are

    known to be exquisitely sensitive to chemical-induced hypoxic

    damage (Nicklas et al., 1992).

    We did not observe increased hindbrain sulfide concentra-

    tions in acutely or subchronically H 2S-exposed animals. Ware-

    nycia et al. (1989) showed that accumulation of sulfide oc-

    curred in the hindbrain of male Sprague-Dawley rats exposed

    to lethal quantities of sodium hydrosulfide (NaHS), an alkaline

    salt that liberates H2S in vivo. Our inability to detect an

    increase in hindbrain sulfide concentrations probably reflected

    the lower (sublethal) doses of H 2S used in our study. We also

    did not observe altered brain cytochrome oxidase activity in

    H2S-exposed animals. Savolainen and coworkers (1980) like-

    wise showed that a single 2-h exposure of mice to 100 ppm

    H2S did not result in inhibition of brain cytochrome oxidase

    activity. We observed increased liver sulfide concentrations in

    rats exposed for 3 h to 200 ppm. Despite the presence of

    elevated liver sulfide concentrations, we did not observe cyto-

    chrome oxidase inhibition in this tissue. Indeed, rats exposed

    to 10 ppm H2S for 3 h had significantly elevated hepatic

    cytochrome oxidase activity. For example, liver cytochrome

    oxidase activity was 136% of preexposure levels following a

    single 3-h exposure to 400 ppm H2S. A similar observation wasnoted by Khan and coworkers (1998), who also observed a

    small increase in liver cytochrome oxidase activity (to 109% of

    control values, not statistically significant) in rats subchroni-

    cally exposed (8 h/day, 5 days/week, for 5 weeks) to 100 ppm

    H2S. The biological significance of this observation is unclear,

    but it implies that respiration in the liver is not inhibited by H 2S

    at these treatment concentrations, and possibly, that H2S de-

    toxification processes in this organ require additional energy

    demands, as reflected by increased respiratory cytochrome

    activity.

    The results of our study provide important new information

    defining the relationship between H2S exposure concentration

    and resulting sulfide concentrations and cytochrome oxidaseactivities in the hindbrain, lung, and nose, each of which is a

    critical target for H2S-induced toxicity. Our results suggest that

    acute exposure to low concentrations ( 30 ppm) of H 2S is

    associated with cytochrome oxidase inhibition in the lung and

    nose. Inhibition of cytochrome oxidase often occurred in the

    absence of elevated tissue sulfide concentration. These data

    suggest that cytochrome oxidase inhibition is a more sensitive

    biomarker of H2S exposure than is tissue sulfide concentration.

    23HYDROGEN SULFIDE TOXICOKINETICS AND TOXICODYNAMICS

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    7/8

    It is not unexpected that measurement of total tissue sulfide

    concentrations is a relatively insensitive biomarker of H2S

    exposure, as most tissues contain high endogenous levels of

    sulfide and this metabolite is highly volatile when unbound.

    More refined studies using radiolabeled H2S with evaluation of

    total and mitochondrial sulfide could better elucidate the dose-

    response relationship between tissue sulfide concentration and

    H2S exposure. Despite the limitations in our experimentaldesign, our data should prove useful in the development of

    biologically based dosimetry and pharmacodynamic models

    for this chemical. The development of dosimetry based models

    should improve the risk assessment for this important environ-

    mental contaminant.

    ACKNOWLEDGMENTS

    This study was funded in part by a grant from the American Petroleum

    Institute (API). The authors thank Drs. Susan Borghoff, Jeffrey Everitt, Greg-

    ory Kedderis, and Barbara Kuyper for their reviews of the manuscript.

    REFERENCES

    ATSDR (1999).Toxicological Profile for Hydrogen Sulfide. Agency for Toxic

    Substances and Disease Registry, U.S. Department of Commerce, Spring-

    field, VA.

    Ammann, H. M. (1986). A new look at physiologic respiratory response to H 2S

    poisoning.J. Hazard. Mater. 13, 369374.

    Arnold, I. M. F., Dufresne, R. M., Alleyne, B. C., and Stuart, P. J. W. (1985).

    Health implication of occupational exposures to hydrogen sulfide. J. Occup.

    Med. 27, 373376.

    Bartholomew, T. C., Powell, G. M., Dodgson, K. S., and Curtis, C. G. (1980).

    Oxidation of sodium sulphide by rat liver, lungs and kidney. Biochem.

    Pharmacol. 29, 24312437.

    Beauchamp, R. O., Jr., Bus, J. S., Popp, J. A., Boreiko, C. J., and Andjelkovich,

    D. A. (1984). A critical review of the literature on hydrogen sulfide toxicity.

    Crit. Rev. Toxicol. 13, 2597.

    Bhambhani, Y. (1999). Acute effects of hydrogen sulfide inhalation in healthy

    men and women. Environ. Epidemiol. Toxicol. 1, 217230.

    Brenneman, K. A., James, R. A., Gross, E. A., and Dorman, D. C. (2000a).

    Olfactory neuron loss in adult male CD rats following subchronic inhalation

    exposure to low levels of hydrogen sulfide. Toxicol. Pathol. 28, 326333.

    Brenneman, K. A., Wong, B. A., Bucellato, M. A., Costa, E. R., Gross, E. A.,

    and Dorman, D. C. (2000b). Direct olfactory transport of inhaled manganese

    (54MnCl2) to the rat brain: Toxicokinetic investigations in a unilateral nasal

    occlusion model. Toxicol. Appl. Pharmacol. 169, 238248.

    Brenneman, K. A., Meleason, D. F., Marshall, M. W., James, R. A., Gross,

    E. A., Martin, J. T., and Dorman, D. C. (in press). Nasal lesions following

    acute inhalation exposure of male CD rats to hydrogen sulfide: Reversibility

    and the possible role of regional metabolic capacity in lesion distribution.

    Toxicol. Pathol.

    Buckley, L. A., Morgan, K. T., Swenberg, J. A., James, R. A., Hamm Jr., J. E.,

    and Barrow, C. S. (1985). The toxicity of dimethylamine in F-344 rats and

    B6C3F1 mice following a 1 year inhalation exposure. Fundam. Appl.

    Toxicol. 5, 341352.

    Cannon, W. C., Blanton, E. F., and McDonald, K. E. (1983). The flow-past

    chamber: An improved nose-only exposure system for rodents. Am. Ind.

    Hyg. Assoc. J. 44, 923928

    Donham, K. J., Knapp, L. W., Monson, R., and Gustafson, K. (1982). Acute

    toxic exposure to gases from liquid manure. J. Occup. Med. 24, 142145.

    Dorman, D. C., Brenneman, K. A., Struve, M. F., Miller, K. L., James, R. A.,

    Marshall, M. W., and Foster, P. M. D. (2000). Fertility and developmental

    neurotoxicity effects of inhaled hydrogen sulfide in Sprague-Dawley rats.

    Neurotoxicol. Teratol. 22, 7184.

    Fischer, L. J., Gracki, J. A., Long, D. T., Wolff, G. T., and Harrison, K. G.

    (2000). Health Effects of Low-Level Hydrogen Sulfide in Ambient Air.

    Report of the Hydrogen Sulfide Investigation Panel, Michigan Environmen-

    tal Science Board.

    Genter, M. B., Owens, D. M., and Deamer, N. J. (1995). Distribution of

    microsomal epoxide hydrolase and glutathione S-transferase in the rat ol-

    factory mucosa: Relevance to distribution of lesions caused by systemically-

    administered olfactory toxicants. Chem. Senses 20, 385392.

    Goodwin, L. R., Francom, D., Dieken, F. P., Taylor, J. D., Warenycia, M. W.,

    Reiffenstein, R. J., and Dowling, G. (1989). Determination of sulfide in

    brain tissue by gas dialysis/ion chromatography: Postmortem studies and

    two case reports. J. Anal. Toxicol. 13, 105109.

    Graedel, T. E., Hawkins, D. T., and Claxton, L. D. (1986). Atmospheric

    Chemical Compounds Sources Occurrence, and Bioassay. Academic Press,

    New York.

    Guidotti, T. L. (1994). Occupational exposure to hydrogen sulfide in the sour

    gas industry: Some unresolved issues.Int. Arch. Occup. Environ. Health. 66,

    153160.

    Hall, A. H., and Rumack, B. H. (1997). Hydrogen sulfide poisoning: Anantidotal role for sodium nitrite? Vet. Hum. Toxicol. 39, 152154.

    Hirsch, A. R., and Zavala, G. (1999). Long-term effects on the olfactory

    system of exposure to hydrogen sulphide. Occup. Environ. Med. 56, 284

    287.

    Hoidal, C. R., Hall, A. H., Robinson, M. D., Kulig, K., and Rumack, B. H.

    (1986). Hydrogen sulfide poisoning from toxic inhalations of roofing asphalt

    fumes.Ann. Emerg. Med. 15, 826 830.

    Jaakkola, J. J., Vilkka, V., Marttila, O., Jappinen, P., and Haahtela, T. (1990).

    The South Karelia Air Pollution Study. The effects of malodorous sulfur

    compounds from pulp mills on respiratory and other symptoms. Am. Rev.

    Respir. Dis. 142, 134450.

    Jiang, X. Z., Buckley, L. A., and Morgan, K. T. (1983). Histopathology of

    toxic responses to chlorine gas in the nasal passages of rats and mice.

    Toxicol. Appl. Pharmacol. 103, 143155.Kage, S., Nagata, T., Kimura, K., and Kudo, K. (1988). Extractive alkylation

    and gas chromatographic analysis of sulfide. J. Forensic Sci. 33, 21722.

    Kage, S., Nagata, T. Takekawa, K., Kimura, K., Kudo, K., and Imamura, T.

    (1992). The usefulness of thiosulfate as an indicator of hydrogen sulfide

    poisoning in forensic toxicological examination: A study with animal ex-

    periments.Jpn. J. Forensic Toxicol. 10, 223227.

    Kangas, J., and Savolainen, H. (1987). Urinary thiosulphate as an indicator of

    exposure to hydrogen sulphide vapour. Clin. Chim. Acta. 164, 710.

    Khan, A. A., Coppock, R. W., Schuler, M. M., and Prior, M. G. (1998).

    Biochemical effects of subchronic repeated exposures to low and moderate

    concentrations of hydrogen sulfide in Fischer 344 rats. Inhal. Toxicol. 11,

    10371044.

    Khan, A. A., Schuler, M. M., Prior, M. G., Yong, S., Coppock, R. W.,

    Florence, L. Z., and Lillie, L. E. (1990). Effects of hydrogen sulfideexposure on lung mitochondrial respiratory chain enzymes in rats. Toxicol.

    Appl. Pharmacol. 103, 482490.

    Kilburn, K. H. (1993). Case report: Profound neurobehavioral deficits in an oil

    field worker overcome by hydrogen sulfide. Am. J. Med. Sci. 306, 301305.

    Kilburn, K. H. (1997). Exposure to reduced sulfur gases impairs neurobehav-

    ioral function. South. Med. J. 90, 9971006.

    Kimura, H. (2000). Hydrogen sulfide induces cyclic AMP and modulates

    NMDA receptor. Biochem. Biophys. Res. Commun. 267, 129133.

    Lopez, A., Prior, M., Lillie, L. E., Gulayets, C., and Atwal, O. S. (1988a).

    24 DORMAN ET AL.

  • 8/10/2019 Toxicol. Sci. 2002 Dorman 18 25

    8/8

    Histologic and ultrastructural alterations in lungs of rats exposed to sub-

    lethal concentrations of hydrogen sulfide. Vet. Pathol. 25, 376384.

    Lopez, A., Prior, M., Yong, S., Lillie, L., and Lefebvre, M. (1988b). Nasal

    lesions in rats exposed to hydrogen sulfide for four hours. Am. J. Vet. Res.

    49,11071111.

    Medinsky, M. A., Kenyon, E. M., Seaton, M. J., and Schlosser, P. M. (1996).

    Mechanistic considerations in benzene physiological model development.

    Environ. Health Perspect. 104(Suppl.), 13991404

    Mitchell, T. W., Savage, J. C., and Gould, D. H. (1993). High-performance

    liquid chromatography detection of sulfide in tissues from sulfide-treated

    mice. J. Appl. Toxicol. 13, 389394.

    Morgan, K. T. (1991). Approaches to the identification and recording of nasal

    lesions in toxicology studies. Toxicol. Pathol. 19, 337351.

    Morgan, K. T., and Monticello, T. M. (1990). Airflow, gas deposition, and

    lesion distribution in the nasal passages. Environ. Health Perspect. 85,

    209218.

    Moulin, F. J. M., Brenneman, K. A., Kimbell, J. S., James, R. A., and Dorman,

    D. C. (in press). Predicted regional flux of hydrogen sulfide correlates with

    distribution of nasal olfactory lesions in rats. Toxicol. Sci.

    NRC (1996). Guide for the Care and Use of Laboratory Animals. National

    Research Council, National Academy Press, Washington, DC.

    Nicholls, P., and Kim, J. K. (1982). Sulphide as an inhibitor and electron donor

    for the cytochrome c oxidase system. Can. J. Biochem. 60, 613623.

    Nicklas, W. J., Saporito, M., Basma, A., Geller, H. M., and Heikkila, R. E.

    (1992). Mitochondrial mechanisms of neurotoxicity. Ann. N.Y Acad. Sci.

    648,2836.

    Reiffenstein, R. J., Hulbert, W. C., and Roth, S. H. (1992). Toxicology of

    hydrogen sulfide. Ann. Rev. Pharmacol. Toxicol. 32, 109134.

    Rickert, D. E., Baker, T. S., Bus, J. S., Barrow, C. S., and Irons, R. D. (1979).

    Benzene disposition in the rat after exposure by inhalation. Toxicol. Appl.

    Pharmacol.49, 417423.

    Rocklin, R. D. and Johnson, E. L. (1983). Determination of cyanide, sulfide,

    iodide, and bromide by ion chromatography with electrochemical detection.

    Anal. Chem. 55, 47.

    Rozman, P., Kim, H. J., Madhu, C., and Klaassen, C. D. (1992). Tissue sulfate

    determination by ion chromatography. J. Chromatogr. 574, 146149.

    Savolainen, H., Tenhunen, R., Elovaara, E., and Tossavainen, A. (1980).Cumulative biochemical effects of repeated subclinical hydrogen sulfide

    intoxication in mouse brain. Int. Arch. Occup. Environ. Health 46, 8792.

    Struve, M. F., Brisbois, J. N., James, R. A., Marshall, M. W., and Dorman,

    D. C. (2001). Behavioral and neurochemical effects of short-term exposure

    to hydrogen sulfide in Sprague Dawley rats. Neurotoxicology 22, 375385.

    Warenycia, M. W., Goodwin, L. R., Benishin, C. G., Reiffenstein, R. J.,

    Francom, D. M., Taylor, J. D., and Dieken, F. P. (1989). Acute hydrogen

    sulfide poisoning. Demonstration of selective uptake of sulfide by the

    brainstem by measurement of brain sulfide levels. Biochem. Pharmacol.38,

    973981.

    Warenycia, M. W., Goodwin, L. R., Francom, D. M., Dieken, F. P., Kombian,

    S. B., and Reiffenstein, R. J. (1990). Dithiothreitol liberates non-acid labile

    sulfide from brain tissue of H2S-poisoned animals. Arch. Toxicol. 64, 650

    655.

    Warneck, P. (1988). Chemistry of the Natural Atmosphere. Academic Press,

    New York.

    Watt, M. M., Watt, S. J., and Seaton, A. (1997). Episode of toxic gas exposure

    in sewer workers. Occup. Environ. Med. 54, 277280.

    Weyant, R. S., Chan, W. C., and Austin, G. E. (1988). Automated marker

    enzyme analysis of density gradient fractions using the Cobas-Biocentrif-

    ugal analyzer. Am. J. Clin. Pathol. 89, 384389.

    25HYDROGEN SULFIDE TOXICOKINETICS AND TOXICODYNAMICS