paudyn et al cold regions 53(1) 102-114

Upload: riswanda-firman

Post on 02-Jun-2018

224 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    1/13

    Remediation of hydrocarbon contaminated soils in the

    Canadian Arctic by landfarming

    Krysta Paudyn a, Allison Rutterb ,, R. Kerry Rowe a, John S. Poland b

    aGeoEngineering Centre at Queen's-RMC, Department of Civil Engineering, Queen's University, Kingston, Ontario, Canada K7L 3N6

    b School of Environmental Studies, Queen's University, Kingston, Ontario, Canada K7L 3N6

    Received 8 December 2006; accepted 24 July 2007

    Abstract

    One of the preferred methods for the remediation of fuel contaminated soil today is landfarming. This is particularly true for

    remote sites because the method requires minimal equipment and is therefore by far the lowest cost option. The term landfarming

    generally refers to the process whereby hydrocarbon contaminated soils are spread out in a layer about half a meter thick, nutrients

    are added, and periodically the soils may be mixed. During landfarming, hydrocarbons can be lost through volatilization or

    bioremediation and thus landfarming refers to the combination of the two processes.

    In the challenging Arctic climate, the performance of landfarming studies has been variable and the relative contribution of the

    two processes has not been studied. This paper describes the successful remediation of diesel-contaminated soils at the former

    military base at Resolution Island, Nunavut. The site is 130 km from the nearest community and this isolation together with very

    inclement weather and average summer temperatures of 3 C presents significant challenges for remediation. Trial landfarm plotswere established in 2003 to compare four sets of conditions; daily aeration, aeration every 4 days, addition of fertilizer with aeration

    every 4 days and a control plot. The field trial has clearly demonstrated enhanced bioremediation when fertilizer was added and

    also significant hydrocarbon losses due to aeration by rototilling. The rate of bioremediation was similar to the rate of volatilization

    in the field trial. In addition to the landfarms established on site, extensive complementary laboratory experiments have been

    carried out. Bioremediation was demonstrated at 5 C in the laboratory reactors and isoprenoid markers indicated increased

    bioremediation with increased temperatures. In the reactor experiments, rate constants for volatilization and bioremediation

    increased with temperature.

    2007 Elsevier B.V. All rights reserved.

    Keywords:Landfarming; Hydrocarbon; Remediation; Arctic; Bioremediation; Volatilization

    1. Introduction

    The term landfarming refers to the process where

    hydrocarbon contaminated soils are spread out in a layer

    of 0.31.0 m thick, nutrients are added and the soils are

    mixed periodically. During the process of landfarming,

    the total petroleum hydrocarbons, (TPH), may be lost

    through volatilization or biodegradation. Landfarming

    refers to the combination of the two processes. Treatment

    regimes for landfarms vary with climate, location, tem-

    perature and soil type. Enhanced bioremediation of con-

    taminated soil typically involves the addition of nutrients

    and water, and periodic tilling to mix and aerate the soil

    (McCarthy et al., 2004). Additional amendments (e.g.,

    Available online at www.sciencedirect.com

    Cold Regions Science and Technology 53 (2008) 102 114www.elsevier.com/locate/coldregions

    Corresponding author. Tel.: +1 613 533 2642; fax: +1 613 533

    2897.

    E-mail address:[email protected](A. Rutter).

    0165-232X/$ - see front matter 2007 Elsevier B.V. All rights reserved.doi:10.1016/j.coldregions.2007.07.006

    mailto:[email protected]://dx.doi.org/10.1016/j.coldregions.2007.07.006http://dx.doi.org/10.1016/j.coldregions.2007.07.006mailto:[email protected]
  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    2/13

    bulking agents to increase aeration (Straube et al., 2003),

    co-substrates to stimulate microbial metabolism

    (Mphekgo and Cloete, 2004) or bacterial inoculations

    (Straube et al., 2003) are sometimes added to speed the

    remediation process. Landfarming has proven consis-

    tently successful in warmer southern climates, (Mc-Carthy et al., 2004). For instance, in the 12-month

    operational period of an Australian landfarm TPH levels

    were remediated from 4644 ppm to less than 100 ppm

    (Line et al., 1996).

    In colder Antarctic and Arctic climates, trials involv-

    ing landfarming or bioremediation have been conducted

    with mixed results (Delille, 2000; Seklemova et al.,

    2001; Aisablie et al., 2004; McCarthy et al., 2004).

    Research has shown the presence of organisms adapted

    to cold conditions at sites where hydrocarbon contam-

    ination is present in these cold climate soils (Mohn andStewart, 2000). Hydrocarbon degrading extremophiles

    are thus ideal candidates for the biological treatment of

    polluted extreme habitats such as the Canadian Arctic,

    (Rike et al., 2003; Mohn and Stewart, 2000). A wide

    variety of microorganisms have been detected in the

    active layer in Arctic soils in northern Canada and

    Alaska (Deming, 2002). These cold habitats possess

    sufficient indigenous microorganisms forin situ biore-

    mediation, (Whyte et al., 1999; Ferguson et al., 2003).

    They adapt rapidly to hydrocarbon contamination in the

    soil, as demonstrated by significantly increased numbers

    of oil degraders shortly after a pollution event (Atlas,1995). An increased number of the hydrocarbon degrad-

    ing bacteria in response to oil spills has been reported by

    bothWhyte et al. (1999) and Rike et al. (2001)illustrat-

    ing that growth and proliferation of hydrocarbon

    degrading bacteria have taken place under site-specific

    conditions. Over the past several years, a number of

    studies in both Arctic and Antarctic regions have shown

    that microorganisms naturally occurring in harsh envi-

    ronments are capable of degrading petroleum hydro-

    carbons (McCarthy et al., 2004; Mphekgo and Cloete,

    2004; Ferguson et al., 2003; Kerry, 1993).Landfarming has been used in cooler locations such

    as alpine environments (Margesin and Schinner, 2001)

    and Alaska (Reynolds et al., 1998) where summer tem-

    peratures are much higher than those of the central and

    eastern Canadian Arctic. Rates of biodegradation and

    volatilization have been shown to be slower at low

    temperatures (Snape et al., 2005), however the relative

    rates and therefore their contributions to landfarming in

    Arctic climates are still relatively unknown. The per-

    centage of remediation attributable to aeration in various

    field studies varies (Reimer et al., 2003; Chatham, 2003;

    Ausma et al., 2002). In order to study the bioremediation

    of hydrocarbons ratios of n-alkanes to pristine and

    phytane can be used for good effect (Atlas, 1995;

    Ferguson et al., 2003). There are many landfarms estab-

    lished in cold climates, however, there is a paucity of

    well documented field trials. This work describes a field

    study specifically designed to assess the relative con-tributions of bioremediation and volatilization using

    aerated and fertilized regimes. The trial landfarm ex-

    periment was established at Resolution Island in the

    summer of 2003 in order to determine the viability of the

    landfarming technology for remediation of TPH con-

    taminated soils at the site. Resolution Island is located at

    the southeastern tip of Baffin Island approximately

    310 km southeast of Iqaluit, Nunavut. The site is

    accessible for fieldwork for approximately 3 months a

    year and during that time often suffers from inclement

    weather; including heavy fog and rainstorms. The aver-age temperature during the summer months is 3 C.

    Resolution Island was the site of one of the military

    bases that formed the Polevault Line. The site has been

    the site of a major cleanup operation as it was highly

    contaminated with polychlorinated biphenyls, (PCBs),

    metals and petroleum hydrocarbons (Poland et al.,

    2001).

    Ex situstudies of landfarming have focused on nut-

    rients (Ferguson et al., 2003; Braddock et al., 1997),

    bioaugmentation (Mohn and Stewart, 2000; Van Veen

    et al., 1997), cold adapted organisms (Kunihiro et al.,

    2005; Mikan et al., 2002), oxygen depletion (Zytneret al., 2001) and water content (Ferguson et al., 2003).

    Many studies attempt to assess the viability of biore-

    mediation using radiolabelled linear hydrocarbons

    (Mohn and Stewart, 2000; Braddock et al., 1997) but

    the extrapolation to field studies is often difficult

    (Zytner et al., 2001). In this study, laboratory reactors

    were designed to model the trial landfarm established at

    Resolution Island and to investigate the effect of

    temperature on volatilization and bioremediation.

    There are many hydrocarbon contaminated sites which

    require remediation in the Arctic and there is currentlyno criterion available to determine if landfarming will be

    viable at a given site. This study outlines the

    experimental design and initial laboratory experiments

    that are intended to establish these criteria and field

    protocols.

    2. Methodology

    2.1. Trial landfarm

    Three truckloads (30 m3) of TPH contaminated soil

    were excavated from two diesel-contaminated areas on

    103K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    3/13

    site. The soil was displaced to an area that had been

    previously leveled to a slope less than 5%. Heavy

    equipment was used to homogenize the soil cache and

    evenly distribute the material to each of the four plots.

    Each plot measured 5 m by 5 m with a depth of

    0.3 m. Maintenance and operation of the triallandfarm included subjecting each plot to a different

    regime (Fig. 1). One plot was established as the

    control plot and had no action other than sampling.

    Aeration was achieved by rototilling. Two plots were

    rototilled only; one every day and the other every

    fourth day. The final plot was rototilled every 4 days

    and, in addition, fertilizer was added to this plot. Thus

    the four plots were maintained as follows: control

    plot; daily aeration; aeration every 4 days; fertilizer

    added with aeration every 4 days. As a result of

    operational experiences during the first season, in thesecond season, the regime of aeration frequency

    considered only fair days; fair days were defined as

    days when it was not raining, snowing or excessively

    foggy.

    The fertilizer was added on day 16. Nitrogen and

    phosphorus were added to the landfarm plots in the form

    of granular agricultural fertilizers. Urea, containing 46%

    nitrogen was the primary source for nitrogen while phos-

    phorus was added as diammonium phosphate, (DAP),

    which contained 20.1% phosphorus and 18.0% nitrogen

    by weight. Nutrient additions were based on applying a

    C:N:P ratio of 100:7.5:0.5. No additional fertilizer wasadded to any of the trial landfarm plots after the initial

    application on day 16.

    The soil at Resolution Island can be characterized as

    follows. It is largely composed of sand and gravel par-

    ticles with only 10% of particles finer than 75 M. Thesoil has no plasticity, the soil pH is 5.8 and the organic

    content is 1.1%.

    2.2. Laboratory reactor design

    The reactor design is presented inFig. 2. The body ofthe individual reactors was constructed with polyvinyl

    chloride (PVC) sewer pipe cut into 0.36 m lengths. The

    length of the tube was chosen to accommodate approxi-

    mately 1.2 kg of soil such that the soil level was below

    the mid-point air inlet of the reactor. The landfarm

    aeration process was simulated by inverting the reactor

    Fig. 1. The trial landfarm at Resolution Island Nunavut. Maintenance regime as indicated in photograph.

    104 K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    4/13

    tube. At the reactor base a cap was glued in place while

    at the top of the reactor tube a removable cap was press

    fitted and sealed with electrical tape. Air was passed

    through the reactors at mid-height through a 0.016-m

    brass fitting which was screwed into the tapped sewer

    pipe and sealed with Teflon tape. For the air outtake, a0.016-m brass fitting, was 0.08 m higher than the inlet

    and at an angle of 120 to it. A ball valve was glued into

    position beside the outtake tube of each reactor to allow

    water to be added to the system. This arrangement

    allowed the inversion of the tube without soil entering

    the air passages. A 2.5-m long, 0.102-m diameter high

    density polyethylene (HDPE) tube was used as an air

    pressure containment vessel between the air source and

    the individual reactors. This tube was affixed to the

    laboratory compressed air line. Holes were drilled and

    tapped into the tubing to support 24 pressure regulatorvalves that were then each attached to a reactor allowing

    regulation of air flow through each reactor. The air

    entering the manifold from the line was controlled to the

    temperature of the room by passing source air through

    10 m of coiled copper pressure tubing. In addition,

    source air was passed through a 0.15-m long charcoal

    filter (0.03 m diameter), which ensured that hydrocar-

    bon contamination from the source line was eliminated.

    Volatilized hydrocarbon was captured at the air outlet of

    each reactor with a granulated activated coconut char-

    coal (GAC) trap. Each trap was 0.20 m long with an

    internal diameter of 0.015 m and filled with approxi-

    mately 20 g of 1240 mesh granulated activated coco-

    nut charcoal.

    2.3. Reactor system

    The individual reactors were designed to reflect thephysical conditions that contaminated soil was subject

    to in a field landfarm. Depth, tillage (inversion of the

    individual reactors), temperature, windspeed, moisture

    content were each considered in the reactor design and

    protocol. By monitoring both volatilized and residual

    hydrocarbon, it was possible to attribute loss of TPH to

    either bioremediation or aeration using a mass balance

    approach.

    Each reactor was filled with 1.2 kg of diesel-con-

    taminated soil from Resolution Island. Air was passed

    through each of the closed vessels and volatilized TPHfrom the soil was collected on charcoal tubes at the

    reactor air outlet. The charcoal filters were replaced and

    analyzed for TPH periodically and the soil in each

    reactor was also analyzed periodically. Moisture content

    was carefully monitored and maintained in the range

    of 1015% by the periodic addition of water. Flow

    rate through each reactor was monitored daily and kept at

    1 L/min. Reactor systems, each comprising 24 reactors,

    were set up in three temperature controlled rooms

    maintained at 5 C, 8 C and 18 C.

    Two sets of reactor experiments are reported on here.

    In the first experimental set, at each temperature, reactors

    Fig. 2. Design of the laboratory reactors used for the ex situwork.

    105K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    5/13

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    6/13

    significance difference between plots on day one of the

    trial but all other samplings showed a significant differ-

    ence between the plots (pb0.05 on day 16,pb0.005 all

    other sampling days). At the end of the first season (day

    31) all three treatment plots were significantly different

    (t-test, p b0.05) than the control plot and the TPH con-

    centration in the fertilizer added, aerated every 4 days

    treatment plot was significantly lower than the other two

    treatment plots (t-test,pb0.05). At the end of the second

    season the same trend continued and it was clear that the

    rate of remediation was as follows: control plotbaeratedevery 4 daysbaerated dailyb fertilized and aerated every

    4 days (t-test, pb0.05). By the end of the third season

    there were no significant differences between the three

    treatment plots but all three treatment plots were still

    significantly different than the control plot (t-test, pb

    0.05). Initial TPH levels for all plots were 2800 ppm and

    final concentration in the fertilizer added plot was below

    200 ppm. Significant variations were noted for TPH

    concentrations in each plot for a given season and the

    average relative standard deviations for the data points in

    Fig. 3 was 28%. Weathered TPH contaminated soil is

    difficult to homogenize which contributes to the problem

    of obtaining consistent data under both field and labo-

    ratory conditions. Lower results were obtained when

    sampling occurred on wet days where the landfarm was

    Fig. 4. C17/Pr Ratios for the trial landfarm over the three seasons.

    Fertilizer was added on day 16. Five samples were taken from eachplot and averaged. Error bars are one standard deviation.

    Fig. 3. Concentrations of TPH in soil for the trial landfarm over the three seasons. Five samples were taken from each plot and averaged. Error bars are

    one standard deviation.

    107K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    7/13

    waterlogged and the sample water content was 15%

    or greater. This is consistent with reports by Fine et al.

    (1997)that sorption of hydrocarbons to soils is reduced

    as water content increases. Despite large variations

    the results show a clear trend. The fertilized plot shows

    a striking decrease in TPH levels by over 90%. Thedaily aerated plot and the aerated every 4 days shows

    losses in excess of 80%. These data are consistent

    with successful landfarm results reported byChatham

    (2003) and Reynolds et al. (1998). The control plot

    TPH levels were reduced (t-test,pb0.05) but remained

    above 1000 ppm after 3 seasons. Rate constants for the

    loss of TPH based on first order reactions are discussed

    below.

    Pristane (Pr), (C19H40) and phytane (Ph), (C20H42)

    are two isoprenoid alkanes present in diesel fuel. Also

    present in diesel fuel are straight chain isomers C17H36(C17) and C18H38 (C18) of similar boiling points to

    pristane and phytane respectively. The branched nature

    of the pristine and phytane compounds makes them

    relatively resistant to biodegradation when compared to

    the linear counterparts (Atlas, 1995). The nature of the

    compounds yields similar rates of volatilization, and

    therefore major volatilization will result in no net change

    in the isomeric ratios. A measurable change in the iso-

    meric ratios is an indication that bioremediation is oc-

    curring in the affected soil (Snape et al., 2005; Kerry,

    1993).

    The C17/Pr ratios were calculated using the chroma-

    tograms of extracted soil samples from the plots to

    determine if significant bioremediation was occurring.

    For the 3 plots with no fertilizer added, (Fig. 4)theC17/Prratios did not change significantly indicating that no

    measurable bioremediation was occurring. Moreover,

    for the plot that had fertilizer applied, the C17/Pr ratios

    were dramatically reduced indicating significant biore-

    mediation had occurred. Similar results were found for

    the C18/Ph ratios although there were small but signi-

    ficant decreases (t-test, pb0.05). for the non-fertilized

    plots. Overall the data therefore demonstrates that biore-

    mediation is viable at the site but shows that aeration can

    also achieve very significant reductions in TPH. The

    addition of fertilizer is low maintenance and economi-cally enticing when comparing this option to daily

    tilling. The fertilized plot was rototilled every 4 days but

    it is not clear how necessary this soil aeration step was in

    assisting the bioremediation; oxygen is necessary for

    bioremediation of TPH to occur. Zytner et al. (2001)

    found evidence of oxygen depletion after 30 days.

    Laboratory experiments described below were set up to

    address this issue and facilitate the development of an

    optimal landfarm operation protocol in cold climates.

    Fig. 5. Concentrations of TPH in soil in the laboratory reactors at 5 and 18 for control, aerated daily, aerated every 4 days and aerated every 4 days andfertilized. Six reactors were used for each regime at each temperature.

    108 K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    8/13

    3.2. Laboratory reactors

    The reactor sets indicated that by aerating every

    4 days and fertilizing, the TPH contaminated soil could

    be remediated at all 3 temperatures. The TPH levels

    generally decreased with time, with the control samples,which were not rotated, showing the least decrease of

    the four regimes. In the first set, after approximately

    5 months, 78% of the TPH had been remediated from

    the soil at 18 C in the fertilized reactors. The TPH in the

    soil at 8 C and 5 C was remediated 36% and 51%

    respectively. For both reactor sets, the decrease in TPH

    was significantly better (pb0.05) at 18 C as compared

    to 5 C, indicating, as would be expected, improved

    remediation with increased temperature. The remedia-

    tion at these lower temperatures is important because it

    demonstrates that this regime is applicable to sites withcolder temperatures than Resolution Island. Cold tempe-

    rature hydrocarbon degraders have been isolated (Rike

    et al., 2001) and remediation at 2 and 5 C has been

    demonstrated by Zytner et al. (2001). The fertilized

    reactors exhibited the largest decrease in TPH con-

    centration over the course of the experiment for all

    temperatures.

    Fig. 5illustrates the differences between the 5 C and

    18 C reactor sets from the first set of reactors. The

    fertilized aerated every 4 days reactor systems were

    significantly different than the control reactor systems

    (t-test, pb0.05) by day 42 at 5 C and by day 54 at18 C. The average relative standard deviation of the

    data points in Fig. 5 is 19%. At 18 C, the results are

    similar to the landfarm with the fertilized and rotated

    every 4 days reactor systems being significantly dif-

    ferent (t-test,pb0.05) than the non-fertilized and rotated

    every 4 days reactor on sampling days 42, 52, 138 and

    169. At 5 C the difference is less clear with aerationalone as effective as fertilizer added and aerated every

    4 days for most of the trial; there is no significantly

    difference (pb0.05) until days 127 and 169 at 5 C. The

    C17/Pr ratios presented inFig. 6indicate that there was a

    C17/Pr isoprenoid ratio decrease in only the reactors to

    which nutrients were added. The decrease in the ratio

    increases with temperature (Fig. 6). The C17/Pr ratios

    indicate that bioremediation is delayed and slower at

    5 C. Results of the second reactor set showed little

    difference between the 4- and 12-day rotation schedule.

    However in the field, daily aeration did result in an

    increased rate of volatilization. For the fertilized set offour reactors that were not rotated but had fertilizer

    added, bioremediation was evident. This result holds

    promise for in situbioremediation.

    Analysis of the charcoal tubes enabled the amount of

    TPH lost through aeration alone to be quantified. Data

    for the first set of experiments is presented inTables 1

    and 2.Table 1indicates the total TPH volatilized from

    the reactors while inTable 2the mass of TPH volatilized

    as a function of time at 18 C is given. As one would

    Table 1

    Total mass of TPH collected on charcoal traps (mg) over duration of

    the experiment (169 days)

    Reactor type Code Total mass of TPH collected on

    charcoal traps (mg)

    18 C 8 C 5 CControl CP 1200 460 390

    Everyday A-1D 2580 1420 1510

    Every 4 days A-4D 2270 1180 1210

    Fertilizer F-4D 1960 1100 1160

    Experimental code: CP = control plot; A = aeration alone while F =

    fertilizer added; -xD = rotation every x days.

    Fig. 6. C17/Pr ratios in the laboratory reactors at three temperature from

    the first reactor set. Six reactors were used for each regime at eachtemperature.

    Table 2Mass of TPH collected during experiment per day (mg/day) at 18 C

    Reactor code Mass of TPH collected during experiment per day

    (mg/day)

    Days from start of experiment

    130

    days

    3151

    days

    5277

    days

    78126

    days

    127169

    days

    CP 9 6 4 10 5

    A-1D 24 17 16 16 8

    A-4D 22 11 11 14 9

    F-4D 21 12 11 10 5

    Experimental code: CP = control plot; A = aeration alone while F =fertilizer added; -xD = rotation every x days.

    109K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    9/13

    expect, the amount of TPH volatilized increased with

    temperature (Ausma et al., 2002). In addition, the mass

    volatilized increased with the frequency of rotating the

    reactors. Thus, for the six control samples averaged in

    the first line ofTable 1, the amount of TPH aerated for

    the three temperatures of 18 C, 8 C, and 5 C was

    1200, 460 and 390 mg respectively. For the 18 C

    reactor sets, the amount of TPH volatilized was 1200 mg

    for the control samples, 2270 mg and 1960 mg for thetwo sets rotated every 4 days and 2580 mg for the set

    rotated every day (Table 1). These results indicate that

    volatilization of TPH increases with a regime that in-

    cludes aeration. Similar trends are seen at all tempera-

    tures in the second set of reactors. The mass of TPH

    volatilized decreased with time (Table 2). This is to be

    expected as the mass of TPH remaining in the soil

    decreases with time as it is removed from the system.

    The charcoal tubes enabled the calculation of a mass

    balance for the reactor systems. In each experiment, no

    net loss or gain of TPH should be evident unlessbioremediation is occurring.Fig. 7indicates the net loss

    of TPH in the second set of reactors. Both sets show that

    TPH was removed from the soil in the fertilized reactors

    by both volatilization and bioremediation while all other

    reactors only exhibit loss through volatilization. These

    experiments indicate that both mechanisms were

    occurring in the fertilized regimes even at temperatures

    as low as 5 C. These laboratory results support the

    results obtained in the field study at Resolution Island. It

    should be pointed out that there is large imprecision for

    TPH concentrations in soil analysis (approximately 30%

    variation) due to the heterogeneity of the soil. This is

    further extenuated when calculating for a mass balance

    and is indicated in a large coefficient of variation for

    each system (Fig. 7).

    3.3. Rate constants

    In previous studies, the TPH degradation has been

    modeled using first order rate constants in both

    unfertilized and fertilized soil systems (Demque et al.,1997; Zytner et al., 2001). For this system, the rate of

    loss of TPH is proportional to its concentration.

    d TPH

    dt kTPH

    where [TPH] is the concentration of TPH and t is the

    time. Integration of above equation yields

    ln TPH t kt ln TPH 0 and TPH t TPH 0ekt

    where [TPH]t is the TPH concentration at time t and

    [TPH]0is the initial concentration of TPH in the system.

    The first order rate constant,k, is the slope of the line for

    a plot of the natural log of [TPH] with respect to time.

    The fit of the line to the average TPH data can be

    described by ther2 value, where a value of 1.0 indicates

    a perfect fit. The fit of the line gives an indication of the

    fraction of the variance in the slope of the line. A

    fraction of variance (r2) of less than 0.90 may be due to

    the large error, associated with the average TPH in soil

    values (20%) used to calculate the first order rate

    constants. The nature of the Arctic climate, especially

    Fig. 7. Net loss of TPH in the second reactor set, {TPH initial (soil)TPH final (soil)1169 TPH (charcoal)}mg=TPHnet. Error bars represent thecoefficient of variation for each regime.

    110 K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    10/13

    the extreme temperatures in the winter season, has led to

    the assumption that biodegradation exhibits a hiatus

    during the winter season when the soil is frozen (Rike

    et al., 2001). However other models (Huesemann and

    Truex, 1996) and some studies (Zimov et al., 1996;Oechel et al., 1997; Jones et al., 1999) have shown that

    bioremediation sometimes occurs in the winter months

    but with significantly reduced microbial activity and in

    no lower than 5 C temperatures (Cline and Schimel,

    1995).

    Rate constants have been calculated for the experi-

    mental landfarm and the laboratory reactors (Tables 3

    and 4). For the landfarm, it was assumed that reme-

    diation did not occur in the winter months (Table 3).

    This is not unreasonable since soil temperatures as low

    as 40 C can be expected from mid-September to late

    June (Schimela et al., 2004). Therefore, the time scalewas adjusted to exclude the winter months, when the

    landfarm was unmanaged, from the rate calculation. For

    the summer field season, the average soil temperature

    was 9.3 C. Rate constants from this experiment agree

    with those from previous studies (Zytner et al., 2001).

    Data for the second laboratory set are presented in

    Table 4; similar results were obtained for reactor set 1.The rate constants have been calculated in two ways

    from the rate of loss of hydrocarbon to the charcoal

    traps, k(GAC) (volatilization only) or from the TPH

    concentration data, k(TPH), (volatilization and biore-

    mediation). Rate constants for bioremediation, k(BIO),

    were obtained by subtraction of the TPH derived rate

    constant aeration rate from the nutrient added equivalent

    treatment.

    An examination of the rate constants with respect to

    the mechanism of aeration indicates that the data can

    generally be described by first order kinetics. In thereactor experiments, rate constants increase with tempe-

    rature and with the frequency of aeration. The k(GAC)

    data yields a high fraction of variance with frequent

    aeration (r2N0.94 in 90% of cases). Without rototilling

    the rate of volatilization is likely dependent on rate of

    diffusion of TPH in soil. The field control plot yields a

    rate constant of 0.007 day1. This predominantly repre-

    sents loss due to aeration and any losses due to erosion.

    For the daily aeration field plot the calculated rate con-

    stant is 0.017 day1 while the 4-day rototilling regime

    yields a value of 0.015 day1. Thus, not unexpectedly,

    the daily aeration regime resulted in a greater loss ofTPH due to volatilization. The C17/Pr ratios for both

    Table 4

    First order rate constants for TPH remediation for reactor set 2

    Temperature Code Rotation frequency k(GAC) r2 k(TPH) r2 k(BIO)

    C day day1 day1 day1

    5 CP 0 0.001 0.89 0.002 0.58

    5 F-0D 0 0.001 0.87 0.006 0.65 0.004

    5 A-4D 4 0.002 0.98 0.002 0.95

    5 F-4D 4 0.002 0.92 0.009 0.67 0.007

    5 A-12D 12 0.001 0.94 0.003 0.50 5 F-12D 12 0.001 0.87 0.009 0.68 0.009

    8 CP 0 0.001 0.90 0.002 0.40

    8 F-0D 0 0.007 0.81 0.007 0.85 0.005

    8 A-4D 4 0.002 0.97 0.005 0.95

    8 F-4D 4 0.002 0.89 0.012 0.83 0.008

    8 A-12D 12 0.001 0.81 0.005 0.78

    8 F-12D 12 0.001 0.84 0.012 0.94 0.007

    18 CP 0 0.002 0.98 0.003 0.83

    18 F-0D 0 0.001 0.88 0.012 0.70 0.009

    18 A-4D 4 0.003 0.97 0.006 0.92

    18 F-4D 4 0.003 0.96 0.016 0.95 0.010

    18 A-12D 12 0.003 0.94 0.005 0.79

    18 F-12D 12 0.002 0.85 0.016 0.67 0.011

    Experimental code: CP = control plot; A = aeration alone while F = fertilizer added; -xD = turning of laboratory reactors every x days.

    Table 3

    Rate constants calculated for the on site landfarm at Resolution Island

    Temperature Code Rototilling frequency k(TPH) r2 k(BIO)

    C day day1 day1

    9 CP 0 0.007 0.58

    9 A-1D 1 0.017 0.80

    9 A-4D 4 0.015 0.94

    9 F-4D 4 0.026 0.83 0.011

    Experimental code: CP = control plot; A = aeration alone while F =

    fertilizer added; -xD = rototilling every x days.

    111K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    11/13

  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    12/13

    in soil in cold climates by landfarming with various

    summer soil temperature conditions as well as contain-

    ment strategies.

    Acknowledgements

    Department of Indian Affairs and Northern De-

    velopment (DIAND) and the Northern Scientific Train-

    ing Program (NSTP), (Indian and Northern Affairs

    Canada).

    References

    Aisablie, J., Balks, J.M., Foght, J.M., Waterhouse, E.J., 2004. Hydro-

    carbon spills on Antarctic soil: effects and management. Envi-

    ronmental Science & Technology 38, 12651274.

    Atlas, R.M., 1995. Bioremediation of petroleum pollutants. Interna-

    tional Biodeterioration and Biodegradation 35, 317327.Ausma, S., Edwards, G.C., Fitzgerald-Hubble, C.R., Halfpenny-Mit-

    chell, L., Gillespie, T.J., Mortimer, W.P., 2002. Volatile hydrocarbon

    emissions from a diesel fuel-contaminated soil bioremediation

    facility. Journal of the Air & Waste Management Association 52,

    769780.

    Braddock, J.F., Catterall, P.H., McCarthy, K., Ruth, M.L., Walworth,

    J.L., 1997. Enhancement and inhibition of microbial activity in

    hydrocarbon-contaminated Arctic soils: implications for nutrient-

    amended bioremediation. Environmental Science and Technology

    31, 20782084.

    Chatham, J.R., 2003. Landfarming on the Alaskan North slope

    historical development and recent applications. 10th Annual Inter-

    national Petroleum Environmental Conference, Houston, TX No-

    vember 1114, 2003. http://ipec.utulsa.edu/Conf2003/Papers/

    chatham_35.pdf.

    Cline, J.S., Schimel, J.P., 1995. Microbial activity of tundra and taiga

    soils at sub-zero temperatures. Soil Biology and Biochemistry 9,

    12311234.

    Delille, D., 2000. Response of Antarctic soil bacterial assemblages to

    contamination by diesel fuel and crude oil. Microbial Ecology 40,

    159168.

    Deming, J.W., 2002. Psychrophiles and polar regions. Current Opinion

    in Microbiology 5, 301309.

    Demque, D.E., Biggar, K.W., Heroux, J.A., 1997. Land treatment

    of diesel contaminated soil. Canadian Geotechnical Journal 34,

    421431.

    Ferguson, S.H., Franzmann, P.D., Revill, A.T., Snape, I., Rayner, J.L.,2003. The effects of nitrogen and water on mineralisation of

    hydrocarbons in diesel-contaminated terrestrial Antarctic soils.

    Cold Regions Science and Technology 37, 197212.

    Fine, P., Graber, E.R., Yaron, B.,1997. Soil interactionswith petroleum

    hydrocarbons: abiotic processes. Soil Technology 10, 133153.

    Rike, A.G., Haugen, K.B., Borresen, M., Engene, B., Kolstad, P.,

    2003. In situ biodegradation of petroleum hydrocarbons in frozen

    arctic soils. Cold Regions Science and Technology 37, 97120.

    Huesemann, M.H., Truex, M.J., 1996. The role of oxygen diffusion in

    passive bioremediation of petroleum contaminated soils. Journal of

    Hazardous Materials 51, 93113.

    Jones, M.H., Fahnestock, J.T., Welker, J.M., 1999. Early and late winter

    CO2 efflux from arctic tundra in the Kuparuk River watershed,

    Alaska, USA. Arctic and Alpine Research 31, 187190.

    Kerry, E., 1993. Bioremediation of experimental petroleum spills on

    mineral soils in the Vestfold Hills. Antarctica. Polar Biology 10,

    423430.

    Kunihiro, N., Haruki, M., Takano, K., Morikawa, M., Kanaya, S.,

    2005. Isolation and characterization of Rhodococcus sp. strains

    TMP2 and T12 that degrade 2,6,10,14-tetramethylpentadecane

    (pristane) at moderately low temperatures. Journal of Biotechnol-ogy 115, 129136.

    Line, M.A., Garland, C.D., Crowley, M., 1996. Evaluation of landfarm

    remediation of hydrocarbon-contaminated soil at the Inveresk

    Railyard, Launceston, Australia. Waste Management 16, 567570.

    Margesin, R., Schinner, F., 2001. Bioremediation (natural attenuation

    and biostimulation) of diesel-oil-contaminated soil in an Alpine

    glacier skiing area. Applied and Environmental Microbiology 67,

    31273133.

    McCarthy, K., Walker, L., Vigoren, L., Bartel, J., 2004. Remediation of

    spilled petroleum hydrocarbons by in situ landfarming at an Arctic

    site. Cold Regions Science and Technology 40, 3139.

    Mikan, C.J., Schimel, J.P., Doyle, A.P., 2002. Temperature controls of

    microbial respiration in arctic tundra soils above and below freez-

    ing. Soil Biology & Biochemistry 34, 17851795.Mohn, W., Stewart, G.R., 2000. Limiting factors for hydrocarbon

    biodegradation at low temperature in Arctic soils. Soil Biology and

    Biochemistry 32, 11611172.

    Mphekgo, P.M., Cloete, T.E., 2004. Bioremediation of petroleum

    hydrocarbons through landfarming: are simplicity and cost-effec-

    tiveness the only advantages? Reviews in Environmental Science

    and Biotechnology 3, 349360.

    Oechel, W.C., Vourlitis, G., Hastings, S.J., 1997. Cold season CO2emission from Arctic soils. Global Biogeochemical Cycles 11,

    163172.

    Poland, J.S., Mitchell, S., Rutter, A., 2001. Remediation of former

    military bases in the Canadian Arctic. Cold Regions Science and

    Technology 32, 93105.

    Reimer, K.J., Colden, M., Francis, P., Mauchan, J., Mohn, W.W.,

    Poland, J.S., 2003. Cold climate bioremediation a comparison

    of various approaches. Proceedings of 3rd Assessment and Reme-

    diation of Contaminated Sites in Arctic and Cold Climates Con-

    ference (ARCSACC), Edmonton, Alberta, pp. 290295.

    Reynolds, C.M., Brayel, A.W., Travis, M.D., Perry, L.B., Iskandar,

    I.K., 1998. United States Army Corps of Engineers, Special Report,

    9720.

    Rike, A.G., Brresen,M., Instanes,A., 2001. Response of cold adapted

    microbial populations in a permafrost profile to hydrocarbon con-

    taminants. Polar Record 37 (202), 239248.

    Schimela, J.P., Bilbrough, C., Welkerc, J.M., 2004. Increased snow depth

    affects microbial activity and nitrogen mineralization in two Arctic

    tundra communities. Soil Biology & Biochemistry 36, 217227.Seklemova, E., Pavlova, A., Kovacheva, K., 2001. Biostimulation-

    based bioremediation of diesel fuel: field demonstration. Biodeg-

    radation 12, 311316.

    Snape, I., Harvey, P.McA., Ferguson, S.H., Rayner, J.L., Revill, A.T.,

    2005. Investigation of evaporation and biodegradation of fuel

    spills in Antarctica: I a chemical approach using GC-FID.

    Chemosphere 61, 14851494.

    Straube, W.L., Nestler, C.C., Hanson, L.D., Ringleberg, D., Prichards,

    P.H., Jones-Meehan, J., 2003. Remediation of polyaromatic hy-

    drocarbons (PAHs) through landfarming with biostimulation and

    bioaugmentation. Acta Biotechnologica 23, 179196.

    Van Veen, J.A., Van Overbeek, L.S., Van Elsas, J.D., 1997. Fate and

    activity of microorganisms introduced into soil. Microbiology and

    Molecular Biology Reviews 61, 121135.

    113K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114

    http://ipec.utulsa.edu/Conf2003/Papers/chatham_35.pdfhttp://ipec.utulsa.edu/Conf2003/Papers/chatham_35.pdfhttp://ipec.utulsa.edu/Conf2003/Papers/chatham_35.pdfhttp://ipec.utulsa.edu/Conf2003/Papers/chatham_35.pdf
  • 8/10/2019 Paudyn Et Al Cold Regions 53(1) 102-114

    13/13

    Whyte, L.G., Bourbonnire, C., Bellrose, C., Greer, C.W., 1999.

    Bioremediation assessment of hydrocarbon contaminated soils from

    the High Arctic. Bioremediation Journal 3, 6979.

    Zimov, S.A., Davidov, S.P., Prosiannikov, Y.V., Semiletov, I.P.,

    Chapin, M.C., Chapin, F.S., 1996. Siberian CO2 efflux in winter

    as a CO2 source and cause of seasonality in atmospheric CO2.

    Climatic Change 33, 111120.

    Zytner, R.G., Salb, A., Brook, T.R., Leunissen, M., Stiver, W.H., 2001.

    Bioremediation of diesel fuel contaminated soil. Canadian Journal

    of Civil Engineering 28 (suppl. 1), 131140.

    114 K. Paudyn et al. / Cold Regions Science and Technology 53 (2008) 102114