You are on page 1of 10

Journal of Archaeological Science 35 (2008) 1453e1462

http://www.elsevier.com/locate/jas

Methods for calculating brine evaporation rates during salt production


D. Glen Akridge*
Arkansas Archeological Society, 5411 W. Wheeler Road, Fayetteville, AR 72704, USA
Received 11 July 2007; received in revised form 17 October 2007; accepted 22 October 2007

Abstract
Salt is recognized by archaeologists as an important commodity due to the biological need for sodium and other cultural uses. Numerous
studies have described the various techniques used in converting brine to crystallized salt, but few, if any, have attempted to quantify the physical
processes of evaporation in pre-industrial societies. Apart from the few areas where salt mining is possible, nearly all forms of salt production
require evaporation of water to concentrate brine and ultimately produce salt crystals. This study quantifies three of the most common evaporation techniques and provides insight into the production rates of salt and fuel requirements. Methods of calculation are provided for determining evaporation through (1) direct solar heating of brine, (2) applied external heat to a vessel, and (3) an immersed heated object (e.g., stone).
These results provide physical constraints on the evaporation process and provide investigators with techniques for estimating efficiency and
total production of prehistoric and historic saltworks.
2007 Elsevier Ltd. All rights reserved.
Keywords: Salt; Evaporation; Brine; Stone boiling; Numerical methods

1. Introduction
Saltmaking from brine has been a common worldwide industry for thousands of years, beginning by at least the fourth
millennium B.C. in Europe (Olivier and Kovacik, 2006) and
by the first millennium B.C. in China (Flad et al., 2005) and
Central America (Andrews, 1983). Solar evaporation of brine
to form salt continues to be a viable commercial process to this
day along coastal areas (Kostick, 2002). The procedures used
in making salt varied by geographic region and resources locally available. The quantity desired by the local population
may have also influenced the choice of salt production
methods. Although the process often involved techniques
such as leaching, extraction, filtering, and burning of saltenriched plants (Adshead, 1992), the final step in salt production invariably required evaporation of water from brine to
precipitate salt crystals. Numerous studies have focused on
the techniques and archaeological remnants of saltworks.
These studies have tended to leave open the question of
* Tel.: 1 479 790 0261; fax: 1 479 575 5453.
E-mail address: dgakridge@hotmail.com
0305-4403/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jas.2007.10.013

efficiency and total salt production, except in the few cases


where historic descriptions of saltworks exist (e.g., Chiang,
1976). Few have attempted to quantify the process of evaporation. Quantifying the process of salt production does more
than provide estimates for the total salt produced. It provides
insights into the required human labor expenditure and the degree to which the local economy depended on salt. The salt
trade would have largely been determined by the amount of
excess salt that could be produced. In the case of brine boiling
to obtain salt, the need for substantial quantities of fuel (e.g.,
wood) often resulted in serious environmental changes to the
surrounding landscape (Early, 1993).
Quantifying the production of salt can be accomplished
through experimentation or by numerical simulation. Experimentation is often advantageous because it can provide
insights into the subtleties of the process that might not be
realized otherwise, but suffers greatly from the need to control
variables. Salt production is a labor intensive and time
consuming task, and in the case of solar evaporation may
take weeks to obtain salt. This makes experimentation difficult, especially in light of the multitude of potential variables
such as ambient temperature and humidity, brine volume and

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

1454

concentration, fire temperature, and vessel heat transfer properties that need to be monitored and maintained. In addition,
experimentation results are generally only relevant for the
tested scenario and provide only limited insight into other untested evaporation conditions. The great utility of numerical
simulations is the speed at which experiments can be performed. Simulations of processes taking hours or weeks can
be performed in seconds allowing for greater exploration of
variable effects and deeper understanding of the physical
mechanisms underlying evaporation.
This paper lays out in detail all the necessary steps in
simulating evaporation in three different scenarios: (1) solar
evaporation, (2) evaporation from an externally heated pan,
and (3) evaporation from a hot immersed object. Each scenario
involves evaporation of a fixed volume of brine to dryness
resulting in the precipitation of salt crystals. For any batch
evaporation the amount of salt produced can be determined by


ms mw 1:52  104 S2 9:50  103 S

where ms is the mass (kg) of salt crystallized, mw is the mass


(kg) of the water evaporated, and S is the initial salt concentration of the brine in wt% NaCl (Fig. 1). For ease of calculation,
the dissolved salt is assumed to be pure sodium chloride. Sodium and chlorine make up 85% of the inorganic constituents
found in seawater (Lide, 1991) and natural inland brines are
often of even higher purity (Bowman, 1956). Consequently,
the presence of other minor components has little impact on
the numerically simulated results. Exceptions may be inland
lakes containing high concentrations of sulfates. In these
cases, adjustments may need to be made to account for the
mass, density, and vapor pressure differences of a sulfaterich brine.
2. Solar evaporation
In solar evaporation a vessel or ponded area containing
brine is allowed to evaporate under the prevailing
20

NaCl Crystallized (kg)

18
16
26.2 wt%

14
12

20 wt%

10
8

15 wt%

6
10 wt%

4
2
0

5 wt%
0

10

15

20

25

30

35

40

45

50

Water Evaporated (kg)


Fig. 1. Graphical representation of Eq. (1). The maximum sodium chloride
brine concentration is 26.2 wt% (at 25  C). For comparison, seawater has
a salt content of about 3.5 wt%.

environmental conditions. This technique works best at low


latitudes where sunlight duration and intensity are highest
and areas with low relative humidity and rainfall. Solar evaporation also becomes the default method when fuel resources
are scarce and boiling of brine is unfeasible. Historically
this technique was common in coastal areas (e.g., LeConte,
1862) and continues to be a viable commercial process worldwide (Kostick, 2002). Solar evaporation could have been practiced at many inland salines where brine concentrations tend to
be high (e.g., Demir and Seyler, 1999) thus reducing total
evaporation time.
2.1. Calculation method for solar evaporation
Calculating evaporation rates for brine solutions requires
a slight modification to the standard Penman (1948) equation
used by hydrologists to determine evaporation from open
water sources. The Penman approach combines the effects
of radiation and aerodynamic forces controlling evaporation
and has been shown to adequately predict evaporation in
a wide variety of environments (e.g., Finch, 2001). The Penman equation is generally expressed as:
lE

D
g
Rn
f u es  e
Dg
Dg

where E is the evaporation rate expressed as mm/day, l is the


latent heat of vaporization, D is the gradient of the vapor pressureetemperature curve, g is the psychometric constant, Rn is
the net solar radiation, f(u) is a function of wind speed, and es
and e are the saturation vapor pressure of water and ambient
water vapor pressure, respectively. In this paper the Penman
equation has been modified to reflect the reduced vapor pressure of a salt solution. A similar approach has been used previously in determining evaporation from saline lakes (Calder
and Neal, 1984). Additional details of the equation and its variables are discussed by Allen et al. (1998) and Shuttleworth
(1993).
2.1.1. Calculating aerodynamic terms
The aerodynamic forces acting on evaporation are primarily
the result of environmental variables governing diffusion of
water molecules away from the liquid surface of water or
brine. The modified Penman equation for predicting potential
evaporation rates from a free water surface requires knowledge
of several local climate variables. The method shown here is
for daily calculation steps and uses 24 h averages for temperature, humidity, and wind speed. For modeling past saltmaking
endeavors, historical weather data including average climate
conditions for each month can be obtained for a variety of locales from several sources, including the National Climatic
Data Center in the United States.
The latent heat of evaporation l (MJ kg1) varies with temperature according to
l 2:501  0:002361T

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

where T is in degrees celsius. The temperature T for daily time


steps is simply defined as the mean of the daily maximum and
minimum temperatures.
T

Tmax Tmin
2

es 0:6108aw exp

17:27T
237:3 T

The activity coefficient of water aw is a function of the concentration of dissolved salts. The following correlation was derived from experimental vapor pressure data published for
sodium chloride (Lide, 1991).
aw 0:0011m2  0:0319m 1

where m is the concentration of sodium chloride in moles per


liter of water. Density of NaCl solutions has been measured by
Romanklw and Chou (1983). Based on their published data,
the following equation was developed to calculate NaCl solution density over the temperature range of 25e45  C and concentration range of 0e26.2 wt%.
 0:012


25
D
2:754  105 S2 6:872  103 S 0:99704
Tsol
7
where D is the solution density (g cm3), S is the NaCl concentration in wt%, and Tsol is the solution temperature ( C).
The saturation vapor pressure is a function of temperature
and the gradient of this function (kPa  C1) is also required
and can be calculated by
4098es
237:3 T

The psychrometric constant (kPa  C1) is given by


g 0:000655P

emax emin
11
2
The vapor pressure e (kPa) can be determined from the relative humidity
e

The saturation vapor pressure es (kPa) is determined from the


average daily temperature and is an indication of the rate at
which water molecules can escape from the liquid surface.
When dissolved salts are present the saturation vapor pressure
is lowered due to the decreased chemical potential of the liquid water. To adjust for this lowering effect the water activity
coefficient aw is inserted into the basic equation.

1455

where P is the atmospheric pressure (kPa). If the atmospheric


pressure is unknown, an approximate value can be calculated
based upon a sites elevation

5:26
293  0:0065z
P 101:3
10
293
where z is height above sea level (m).
In a similar fashion to the mean temperature, the daily
mean vapor pressure is the average of the maximum and minimum values.

Hr es
100

12

where Hr is the relative humidity (%). Wind speed is incorporated using empirically determined coefficients for the atmospheric resistance encountered in diffusion of the water
vapor away from a liquid surface. For an open water surface
the wind function is given by
f u 6:431 0:536U2

13

where the wind speed U2 (m s1) is measured at 2 m above the


surface.
2.1.2. Determining net radiation
Solar radiation aids in promoting evaporation by imparting energy into the absorbing material. The radiational
energy available at the ground surface is a combination of
both short and long-wavelength radiation and is the difference between the upward and downward radiation fluxes.
The amount of solar energy reaching the ground surface
can be reduced by cloudiness and atmospheric interferences
or increased with increasing altitude. This net radiation, Rn,
can be determined in three ways: (1) direct measurement
using a radiometer, (2) published tables based on latitude,
or (3) calculations incorporating Earths orbital characteristics. Radiometers measure the net radiation by monitoring
the temperature difference across two parallel plates. They
require periodic calibration and each measurement locale
must be generally free of any obstructions blocking incoming or outgoing radiation. Approximate values for the net
radiation reaching the ground surface can also be found in
published tables (e.g., Lide, 1991). These tables are generally organized by latitude and indicate the average net
radiation for a cloudless sky during each month of the
year. These tables do not usually account for altitude
differences but are generally accurate to within 10% during
summer months and 15% during winter months (Lide, 1991).
2.1.3. Calculating net radiation
The following series of equations are required when determining the net radiation by accounting for Earths orbital characteristics. The extraterrestrial radiation Ra (MJ m2 day1)
can be calculated for any latitude and day of year by adjusting
the solar constant Gsc for the solar declination
Ra

2460
Gsc dr us sin4sind cos4cosdsinus 
p
14

where Gsc 0.0820 MJ m2 min1, dr is the inverse relative


EartheSun distance, us is the sunset hour angle (rad), 4 is

1456

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

the latitude (rad), and d is the solar declination (rad). Angles in


radians can be obtained by converting decimal degrees.
p
Decimal degrees
15
180
The inverse relative EartheSun distance dr is given by


2p
J
16
dr 1 0:033 cos
365

Radians

where J is the number day of the year (e.g., 1 for January 1 or


365 for December 31). The solar declination d is


2p
d 0:409 sin
J  1:39
17
365
The sunset hour angle us is given by
us arccos tan4tand

18

The number of daylight hours N can be determined by


24
us
19
p
The solar radiation is calculated by adjusting the extraterrestrial radiation for the relative sunshine duration

n
20
Rs as bs Ra
N

where Rs is the solar radiation (MJ m2 day1), n is the actual


duration of sunshine (hours), N is the maximum possible
amount of sunshine (hours), and as and bs are regression parameters with recommended values of 0.25 and 0.50, respectively. The effective solar radiation reaching the ground
surface is reduced by solar reflection caused by albedo
Rns 1  aRs

21

where Rns is the net solar radiation (MJ m2 day1), and a is
the albedo of the surface. For open water Shuttleworth
(1993) recommends an albedo value of 0.08. However, for
shallow evaporation pans the underlying reflectivity of the
pan must be considered. For example, dark earthenware or
wooden pans filled with water may have an albedo closer to
0.05. When evaporating brine, salt crystals will begin to
form, thus increasing the reflectivity. Rife et al. (2002) found
that albedo values of 0.3 were appropriate for modeling diurnal weather cycles over a salt-encrusted playa.
The flux of long-wave radiation reflected by the ground
back into space is given by the StefaneBoltzmann law minus
that which is absorbed by clouds, water vapor, dust, and carbon dioxide. The net long-wave radiation can be determined
by
 4



4

p
T Tmin
Rs
Rnl s max
0:340:14 e 1:35
0:35
2
as bs Ra
22
where s is the StefaneBoltzmann constant (4.903 
109 MJ K4 m2 day1), T is the maximum and minimum

temperature during a 24 h period (K), e is the water vapor


pressure (kPa), as and bs are regression terms mentioned earlier, Rs is the solar radiation (MJ m2 day1), and Ra is the extraterrestrial radiation (MJ m2 day1). The net radiation Rn
(MJ m2 day1) is simply the difference between the incoming
effective solar radiation and outgoing long-wave radiation.
Rn Rns  Rnl

23

2.2. Results from numerical simulations


To illustrate the utility of predicting solar evaporation rates,
two examples from the literature will be briefly discussed here.
These examples were selected because sufficient evaporation
variables are specified to allow for numerical modeling of
the described saltmaking process. The first example is an ethnographic description of saltmaking from the western coast of
Mexico. The second simulation uses historic descriptions of
19th century Chinese coastal tradition using portable wooden
pans to evaporate brine. Each case requires the input of numerous weather data and site specific information. Historical
weather data including long-term averages can be obtained
from a variety of sources. The data used here were derived
from the online databases of the National Climatic Data Center of the National Oceanic and Atmospheric Administration
(www.noaa.gov); and Weather Underground (www.wunderground.com). Model inputs are summarized in Table 1.
Williams (2002) describes a traditional Mesoamerican saltworks operating on the western coast of Mexico in the state of
Michoacan. The La Placita saltworks operates during the dry
season, roughly from April to mid-June. Specialists, called salineros, scrape the upper surface of soil from a nearby dry estuary. The salt-encrusted sand called salitre is first filtered
through a tapeixtle using seawater and the resulting enriched
brine is transferred to specially prepared evaporation pans
called eras. According to Williams, La Placita has 18 eras
with most measuring approximately 3 m  3 m in dimension.
Each era is lined with beach sand and lime and can hold about
400 L of brine. Every day 2e3 buckets of brine must be added
to maintain a consistent level. It takes 5 days of evaporation to
concentrate the brine sufficiently to begin collecting salt, about
25e30 kg are then collected every other day from each era.
An average of 7 tons of salt can be produced at each saltmaking site during the dry season.
Table 1
Monthly weather averages for selected saltmaking locales

Latitude
Month
Temperature ( C)
Humidity (%RH)
Solar Radiation (MJ m2 day-1)
Wind Speed (m s1)

Michoacan, Mexico

Shanghai, China

16.83 N
May 2000
27.2
77
29.1
5

31.17 N
July
28.4
83
30.8
4

Weather data were obtained from the National Climatic Data Center of the
U.S. National Oceanic and Atmospheric Administration (www.noaa.gov). Solar radiation for a cloudless day was calculated from Section 2.1.3.

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

Based on the information that Williams (2002) supplies, an


estimate for the average evaporation rate can be derived. During the first 5 days of evaporation, the brine concentrates up to
a maximum value of about 26.2 wt%. Once the maximum
value is reached, further evaporation results in the formation
of salt crystals. Eq. (1) indicates that producing 12.5e15 kg
of salt per day requires the evaporation of 35e43 kg of water.
Converting the mass of the water evaporated to volume and dividing by the surface area of an era (w9 m2) results in an
evaporation rate of 3.9e4.8 mm day1.
Using the steps outlined in Section 2.1 we can numerically
simulate the evaporation process during Williams field season
at La Placita in 2000. Input values used for the model are listed
in Table 1. The albedo of the pan is expected to be relatively
high (w0.3) owing to the lime-lined pan and the constant precipitation of salt crystals at the bottom of the pan. Results for
the numerical simulation suggest that the evaporation rate
should be in the range of 4.1e4.7 mm day1 following the
initial 5-day concentration period, essentially identical to that
obtained in practice. The range in evaporation rate given by
the numerically calculated results reflects the potential variation in cloud cover from 0 to 20%. Assuming an initial brine
concentration of about 15 wt% and an average cloud cover of
10%, Fig. 2 represents model results from the numerically simulated evaporation occurring from a single era at La Placita during the first 20 days of May 2000. Interestingly, extrapolating
the production rate of the era in Fig. 2 to the entire field season
(w75 days) yields approximately 790 kg of salt. This correlates
to an average of about 9 eras in use at La Placita based on the
estimate of 7 tons of salt produced each season. This agrees
well with Williams (2002: p. 243) assertion that not all of
the 18 eras are in use at one time.
Similarly, solar evaporation along the coastal margin of
China utilized enriched brine obtained by leaching salty earth
or in some cases from ashes that were spread onto the ground
(Chiang, 1976). The practice of using portable wooden pans
for evaporation began in the 18th century in the Hangchou
Bay area. Although variations in size were common, the typical wooden pan was about 2.5 m long, 1 m wide, and 3e
6 cm deep. Chiang (1976: p. 526) states that crystallization

200
Evaporation Rate, mm/day
Salt Produced, kg

180
160
140

120
100

80
60

Salt Produced (kg)

Evaporation Rate (mm/day)

40
20

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Day of Month (May 2000)


Fig. 2. Numerical simulation results for a 9 m2 era at La Placita saltworks in
Michoacan, Mexico.

1457

of salt in these portable pans took only 1 day under fine


weather in summers. Although we are not told of the brine
strength or volume in these pans, some reasonable assumptions can be made. If the brine is near saturation (26.2 wt%)
and averages 3 cm in depth, then numerical calculations
indicate that the evaporation rate would be 4.8 mm day1
(see Table 1 for model inputs). Upon complete evaporation
the depth of the salt crystals would be about 5 mm indicating
the total evaporation time to be about 5 days. This is somewhat
longer than the estimate given by Chiang (1976) unless he was
referring to incipient crystallization which could begin on the
first day.
The two examples provided here illustrate the value of
numerical modeling in addition to historic descriptions. Computer simulation of the well documented La Placita saltworks
in Mexico provides remarkable agreement between the
authors observations and numerical results. However, historic
descriptions of 18th century Chinese saltmaking appear to
differ from the numerically calculated evaporation rate.
Numerical modeling provides an independent means for validating historical claims regarding salt production.

3. Evaporation from an externally heated pan


This technique typically involves the suspension of a vessel
over a fire or the emplacement of a vessel directly onto a bed
of hot coals. Heat is transferred through the base and walls of
the vessel and warms the interior fluid. The amount of heat
transferred to the brine is governed by the energy output of
the fire and efficiency of heat transfer in a particular brineboiling setup. The efficiency is defined as the amount of
heat transferred into the brine relative to the total heat produced by the fire. For an open fire efficiency is low owing
to the loss of heat by combustion gases escaping around the
vessel. Glanville (2005) calculated an efficiency of about
20% for the 19th century methods of brine boiling in iron vessels described by LeConte (1862) along the east coast of the
United States. This value is somewhat better than the approximately 15% efficiency for boiling water in cooking vessels
over an open fire (McCracken and Smith, 1998) but less
than the more optimal 28e40% efficiency of well vented
wood stoves (Joshi et al., 1991). For earthenware vessels
with substantially lower thermal conductivity than iron
(Table 2), efficiency is reduced even further and may be as
low as 2e5% (Glanville, 2005).
Although efficiency is a poorly known variable and is
unique to each setup for brine boiling with externally applied
heat, the amount of salt produced can be determined by simply
estimating the temperature on the exterior surface of the vessel. This method eliminates (or sidesteps) the issue of efficiency and focuses directly on the heat being transferred
through the vessel and into the brine. Thus, the amount of
heat calculated would be the absolute minimum required to accomplish evaporation and would represent a system of 100%
heat transfer efficiency. Heat loss by escaping combustion
gases, evaporating water molecules, and conduction through

1458

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

Table 2
Physical properties of various materials at 20  C
Material

Density (g cm3)

Thermal conductivity (W m1 K1)

Specific Heat (J g1 K1)

Source

Clay (9.3e18.3 wt% H2O)


Brick, clay
Iron (100%Fe)
Cast iron (3.16%C)
Copper
Bronze (89%Cu 11%Sn)
Lead

1.33e1.50
1.60e1.82
7.87
7.15
8.96
8.77
11.36

0.4e0.7
0.41e0.63
74.48
46.9
393.7
70.6
34.7

0.92
0.4473
0.837
0.3846
0.3771
0.1287

C
A, B, F
D
D, G
D
E
D

Data from [A] Dondi et al. (2004); [B] Bhattacharjee and Krishnamoorthy (2004); [C] Abu-Hamdeh and Reeder (2000); [D] Davis (1998); [E] Copper Development Association (2007); [F] Lange (1952); [G] MatWeb (2007).

vessel walls to surrounding air all contribute to lowering the


overall efficiency.

q
l

26

using Eq. (3) to determine the latent heat of vaporization.


3.1. Calculation method for evaporation by externally
applied heat

3.2. Numerical simulation of applied external heat

By concentrating solely on the heat transferred through


a vessels wall, a one-dimensional steady-state heat transfer
equation can be applied
k
q A Te  Ti
Dx

24

where q is the heat transfer rate in joules per second (J/s), A is


the heated surface area, Dx is vessel thickness, k is the thermal
conductivity in watts per meter-kelvin (W m1 K1), and Te
and Ti are the temperature (K) of the vessel exterior and interior surfaces, respectively (Geankoplis, 2003). This equation
assumes that a constant flow of heat is applied to the vessel exterior. Materials of high thermal conductivity (Table 2) allow
for substantially more heat to be transferred through the vessel
whereas increasing the vessels wall or basal thickness reduces
heat flow. For most practitioners of brine boiling, the primary
means of evaporation control comes from varying the amount
of external heat applied to the vessel which in turn determines
the temperature of the vessels exterior. The temperature of the
internal surface Ti is typically the same as the boiling point of
the brine and will change as evaporation proceeds and the
brine concentrates to a maximum value of about 29.0 wt%
at the boiling point. Although at very high external temperatures, the interior surface may exceed that of the boiling brine.
The boiling point of sodium chloride brine can be estimated by
the boiling point elevation equation
Tb Kb mi 100  C

25

where Tb is the boiling temperature ( C) of the brine, Kb is the


proportionality constant for water (0.512  C/m), m is the brine
concentration (mol NaCl per L of water), and i is the number
of dissociated ions per formula unit (NaCl 2). At high altitudes where pure water boils at temperatures below 100  C,
the appropriate boiling point should instead be inserted into
the equation. Finally, the evaporation rate (g/s) can be determined by

Simulating the effects of brine boiling requires knowledge


of both the physical characteristics of the pan and the amount
of applied heat (i.e. temperature on pan exterior). Often one or
the other of these criteria is unknown for a specific saltwork.
The discovery of several late Roman age lead pans in Britain
provides constraints on these variables which can be used to
demonstrate the numerical model. Lead melts at 327  C and
salt practitioners would likely have kept the external temperature well below this value to avoid accidental melting of the
pan. At Shavington, Cheshire, a lead pan measuring 100 cm
by 90 cm by 14 cm deep was found cut into eight pieces, presumably as a first step to recycling the material (Penney and
Shotter, 1996). At 0.8 cm thickness, the original weight of
the pan would have been approximately 118 kg. Although
none of the surviving Roman era lead pans have been found
in situ, they likely would have been placed across earthen
flue trenches much like many ceramic salt pans elsewhere in
Britain (e.g., Bradley, 1992). Fires were presumably kept relatively low, 200e250  C, to prevent damage to the pan.
Numerically simulating the evaporation from the Shavington salt pan requires only an estimation of the initial brine concentration. Here an arbitrary 10 wt% is used, but this value has
only a minor impact on the duration of the evaporation process. Even if the temperature on the pans exterior is kept relatively low (w200  C), Fig. 3 demonstrates a fairly quick
evaporation for the Shavington pan. Neglecting substantial
heat loss, the brine is brought to a boil in only 3 min and continues until 14 min, whereby evaporation has resulted in the
maximum allowable brine concentration. Further evaporation
after 14 min results in salt crystals forming inside the pan. Ultimately, about 14 kg of salt can be recovered in less than
20 min of boiling, assuming 10 wt% brine in a 126 L pan.
4. Evaporation using hot immersed objects
A third evaporation scenario considered here is that of a hot
object (e.g., stone) placed inside a pan of brine. This method is
believed to have been utilized for salt production in eastern

35

20

30

Brine Concentration 18
16

25

Crystallized Salt

14
12

20

10
15

8
6

10

4
5
0

Crystallized Salt (kg)

Brine Concentration (wt% NaCl)

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

2
0

10

15

20

25

0
30

1459

analytical solution to the conduction equation becomes considerably more complex.


Finite difference methods can be used to obtain a numerical
solution to Eq. (27), which involves computing temperature
changes after small positional (n) and time (t) steps are incremented (Geankoplis, 2003: pp. 386e387; Carslaw and Jaeger,
1959). The positional steps effectively divide the object into
concentric rings that are Dx (m) thick. For a sphere, the temperature can be determined by

1 2n 1
2n  1
T n1 M  2t T n
T n1
29
tDt T n
M 2n t
2n t

Time (min)
Fig. 3. Brine evaporation from a lead pan placed over fire. Pan conditions are:
pan exterior 200  C, pan volume 126 L, pan thickness 0.8 cm, and pan
heated area 0.9 m2. Initial pan heat-up time is negligible.

where M Dx2/(aDt). At the center (n 0) the equation


changes to
tDt T 0

North America from about A.D. 1000e1400 (Brown, 1980,


1981). Stone boiling as a cooking technique probably began
with the introduction of pottery (ca. 2500 B.C.), if not earlier,
and continued into the historic period (Sassaman and Rudolphi, 2001). Thick-walled ceramic pans, most in association
with salines, have been found with capacities ranging from 40
to 400 L. The enormous size and weight of these vessels when
filled with brine would have made them practically immovable
and suspension over a fire seems equally unlikely. Brown
(1980, 1981) concluded that these salt pans were placed in basin-shaped ground depressions and heated stones from nearby
fires were dropped into the pan to facilitate evaporation. The
lack of exterior discoloration from fires on many pans and
the occasional find of stones inside pans (e.g., Bushnell,
1907) lend support to the conclusion that stone boiling was
at least one method utilized by Native Americans to evaporate
brine.
4.1. Calculation method for stone boiling
A hot object transfers its internal heat energy to the surrounding fluid through convective heat transfer at the objects
surface. This transfer of heat is governed by the unsteady-state
conduction equation
vT
v2 T
a 2
vt
vx

27

where a is the thermal diffusivity (m2/s), t is time (s), and T is


the temperature (K) of the object. Thermal diffusivity is
simply
k
a
rCp

28

where k is the thermal conductivity (W m1 K1), r the density (kg m3), and Cp the heat capacity of the object
(J g1 K1). For simplicity, here it is assumed that the object
is spherical and that only the one-dimensional  direction
need be considered. For irregularly shaped objects, the

4
M4
T1
T0
Mt
M t

30

where M  4 for both Eqs. (29) and (30). At the surface, equations accounting for convection must be utilized assuming that
the heat capacity of the outer half-slab can be neglected
tDt T n

nN
2n  1=2
tDt T a 2n  1
tDt T n1
2n  1
nN
nN
2
2

31

where Tn represents surface temperature and Tn1 the temperature at 1 positional step below the surface.
4.2. Numerical simulation of stone boiling
Solving these equations allows for the determination of the
average stone temperature with time after immersion. Initially,
the heat transferred raises the temperature of the brine up to its
boiling point. Any additional heat released by the stone serves
to evaporate water and concentrate brine. Eventually the brine
is concentrated to a maximum value of about 29.0 wt% at its
boiling point with further evaporation resulting in the formation of salt crystals.
For the numerical model, the boiling stone is assumed to be
chert. Chert was well known to Native Americans and was
commonly used to make stone tools. Although the physical
properties of chert are not well studied, there are numerous
studies on the analogous material of amorphous or fused
quartz. The thermal conductivity of chert can be determined
from a polynomial fit of published data by Kanamori et al.
(1968) and Clauser and Huenges (1995)
k 4:67  109 T 3  8:53  106 T 2 6:29  103 T
 9:85  102

32

where k is thermal conductivity (W m1 K1), and T is the


stone temperature (K). Similarly, Robie et al. (1978) have developed expressions for the heat capacity of quartz
Cp 44:603 3:7754  102 T  1:0018  106 T 2

33

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

where Cp is the heat capacity (J mol1 K1) and T is the temperature (K) for the range of 298e844 K. At higher temperatures the equation changes to
Cp 58:928 1:0031  102 T

34
3

20

0.50

18

0.45

16

0.40

14

0.35

12

0.30

10

0.25

0.20

0.15

0.10

Heat transferred, MJ
Salt Produced, kg

2
0

Salt Produced (kg)

Heat Transferred (MJ)

for the range 844e1800 K. The density of chert is 2.6 g cm .


This information together can be used to determine the thermal diffusivity (a) properties of chert for the numerical model.
Fig. 4 graphically represents a simulated evaporation of
placing 700  C chert stone(s), representing 25% of the total
40 L volume, into a salt pan containing 10 wt% brine at
25  C. Boiling begins immediately for brine surrounding the
stone and continues for about 4 min, by which time the stone
temperature now equals that of the brine and no further heat is
transferred. During the first 2 min the evaporating brine concentrates from 10 wt% up to 29.0 wt%. At 1.6 min salt crystals
begin forming and continue until the stone and brine reach
thermal equilibrium. For this scenario, 373 g of crystallized
salt would be produced. Emplacement of additional hot stones
into the pan would continue the evaporation process.
From the short timescales involved to obtain salt, it seems
this method would clearly be effective in evaporating brine.
Unlike suspension of a pan over a fire, stone boiling releases
all of its internal heat directly into the brine. However, there
may be practical limitations for manipulating large volumes
of very hot stones. Stones heated to high temperatures often
shatter and large stones would be difficult to transport from
a fire to the brine pan. Smaller stones would make handling
easier, but would require more repeated firings to achieve
the same evaporation as large stones. The smallest salt pan
found at the Kimmswick site near St. Louis (Bushnell, 1907)
had a volume of approximately 40 L. Assuming a scenario
similar to that outlined above, emplacement of 25 vol% stone
into the salt pan would translate to 26 kg of extremely hot
stone(s) that would have been manipulated. This seems to suggest that stone boiling may actually require a tremendous
amount of human labor to achieve significant quantities of
salt. Fig. 5 indicates the potential evaporation that could be

0.05
9

0.00

Time (minutes)
Fig. 4. Numerical simulation results for Native American stone boiling. Represented here are results for placing a 700  C stone with a volume of 10 L into
a ceramic pan containing 30 L of 10 wt% brine initially at 25  C. The stone is
assumed to be a spherical nodule of chert.

25

Water Evaporated (kg)

1460

20

Volume Ratio
Brine/Stone
1.0
2.0
3.0
4.0

15

10

0
200

300

400

500

600

700

800

900

1000

Initial Stone Temperature (C)


Fig. 5. Evaporation curves for stone boiling with various brine:stone volume
ratios. Curves represent an initial brine temperature of 25  C and 10 wt%
concentration.

obtained with various brine/stone volume ratios. The 10 wt%


brine is assumed to initially be at 25  C before emplacement
of a hot spherical stone of chert. Even for a Vb/Vs 1 and
an initial stone temperature of 1000  C, the mass of the stone
would still exceed the mass of the evaporated water by a factor
of 2.5. Lower stone temperatures or smaller volumes of stone
would result in greater discrepancies between stone mass and
evaporated water mass.
4.3. Hot objects with negligible internal resistance
An alternative mathematical treatment can be applied when
the hot object has negligible internal resistance to the flow of
heat. These objects (e.g., copper and iron) have high thermal
conductivities and will have an approximately uniform internal temperature profile at any given time following immersion
into brine. To maintain an energy balance, heat loss through
the vessel wall or to the atmosphere is assumed to be minimal
relative to the rapid rise in brine temperature. The total amount
of heat transferred at any given time can be determined by


Q Cp rVT0  TN 1  ehA=Cp rVt
35
where Q (J) is the total heat transferred, V (m3) is the volume
of the object, h is a heat transfer coefficient for natural convection (w5000 W m2 K), A is the heated area (m2), and t is the
elapsed time (s) after immersion (Geankoplis, 2003: pp. 277e
286, 357e359).
5. Conclusions
The goal of this paper is to develop a mathematical basis
for brine evaporation that can be used by investigators to quantitatively describe salt production activities. The methods described herein cover three distinct techniques for evaporating
brine: (1) solar evaporation, (2) boiling due to an externally
applied heat source, and (3) boiling caused by a hot immersed
object. Historically these techniques were sometimes used in
combination to achieve the desired evaporation results. Input

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

parameters for the three models range from common weather


variables for solar evaporation to pan and stone physical
dimensions for brine boiling. No one evaporation method
can be considered superior since other factors such as fuel
availability, sunshine intensity, brine concentration, and even
the cultural value placed on human labor make the choice of
evaporation technique unique to each culture. However, the
calculation methods given here do allow for direct comparisons to be made regarding evaporation efficiency. Brine boiling offers the quickest means of evaporation, typically on
the order of minutes for the examples given here, but this
does not account for the time spent acquiring fuel. Solar evaporation does not require fuel but may take days or weeks to
accomplish and is limited to geographic areas with high evaporation and little precipitation.
Although beyond the scope of the current paper, quantifying salt production opens up new areas of research that can
be addressed. Changes in production methods through time
can be the result of either cultural or technological factors,
or both. In a simple least-cost model, humans would be expected to seek strategies to minimize labor and fuel requirements in the salt production process. Salt production has
often been described as a labor intensive endeavor requiring,
in the case of brine boiling, extraction of substantial quantities
of fuel from the local environment. The calculations presented
herein offer a direct means to evaluate technological changes
for improved evaporation efficiency. For example, Eq. (24)
describes the direct relationship between vessel thickness
and the rate of heat transfer. Thinner vessel walls require
proportionately less fuel to achieve evaporation. In addition,
determinations of minimum energy (fuel) requirements can
be made which provide insight into local deforestation or other
resource depletion. Early (1993) noted that the removal of fruit
bearing plants and nut bearing trees from the vicinity of a saltworks may have altered diet and even changed the local population of wild animals that depend on these resources. These
and other concerns can now be better addressed by quantifying
the brine evaporation process.
Acknowledgments
The author thanks Ann M. Early and Dan F. Morse for
helpful discussions on this subject. Ashley Dumas and Robert
C. Mainfort provided insightful comments on an early draft of
this paper. Valuable comments were also received from two
anonymous reviewers that greatly improved this manuscript.
References
Abu-Hamdeh, N.H., Reeder, R.C., 2000. Soil thermal conductivity: effects of
density, moisture, salt concentration, and organic matter. Soil Science Society of America Journal 64, 1285e1290.
Adshead, S.A.M., 1992. Salt and Civilization. St. Martins Press, New York.
Allen, R.G., Pereira, L.S., Raes, D., Smith, M., 1998. Crop evapotranspiration:
guidelines for computing crop water requirements. In: Irrigation and
Drainage Paper, 56. Food and Agriculture Organization of the United
Nations, Rome.

1461

Andrews, A., 1983. Maya Salt Production and Trade. University of Arizona
Press, Tucson.
Bhattacharjee, B., Krishnamoorthy, S., 2004. Permeable porosity and thermal
conductivity of construction materials. Journal of Materials in Civil Engineering 16, 322e330.
Bowman, H.H.M., 1956. Salinity data on marine and inland waters and plant
distribution. The Ohio Journal of Science 56 (2), 101e106.
Bradley, R., 1992. Roman salt production in Chichester Harbour: rescue excavations at Chidham, West Sussex. Britannia 23, 27e44.
Brown, I.W., 1980. Salt and the Eastern North American Indian: an archaeological study. In: Lower Mississippi Survey Bulletin, No. 6. Peabody Museum, Harvard University.
Brown, I.W., 1981. The Role of Salt in Eastern North American Prehistory.
Louisiana Archaeological Survey and Antiquities Commission Anthropological Study No. 3. Department of Culture, Recreation and Tourism, Baton Rouge.
Bushnell, D.I., 1907. Primitive salt-making in the Mississippi Valley, I. Man
No. 13. London.
Calder, I.R., Neal, C., 1984. Evaporation from saline lakes: a combination
equation approach. Hydrological Sciences Journal 29, 89e97.
Carslaw, H.S., Jaeger, J.C., 1959. Conduction of Heat in Solids, second ed.
Clarendon Press, Oxford.
Chiang, T.C., 1976. The production of salt in China, 1644e1911. Annals of
the Association of American Geographers 66, 516e530.
Clauser, C., Huenges, E., 1995. Thermal conductivity of rocks and minerals.
In: Ahrens, T.J. (Ed.), Rock Physics and Phase Relations e a Handbook
of Physical Constants. AGU Reference Shelf, vol. 3. American Geophysical Union, Washington, pp. 105e126.
Copper Development Association. <http://www.copper.org> (accessed
03.07.07.).
Davis, J.R. (Ed.), 1998. Metals Handbook. ASM International, Materials Park,
Ohio.
Demir, I., Seyler, B., 1999. Chemical composition and geologic history of saline waters in Aux Vases and Cypress Formations, Illinois Basin. Aquatic
Geochemistry 5, 281e311.
Dondi, M., Mazzanti, F., Principi, P., Raimondo, M., Zanarini, G., 2004. Thermal conductivity of clay bricks. Journal of Materials in Civil Engineering
16, 8e14.
Early, A.M., 1993. Caddoan saltmakers in the Ouachita Valley: the Hardman
Site. In: Research Series, No. 43. Arkansas Archeological Survey,
Fayetteville.
Finch, J.W., 2001. A comparison between measured and modeled open water
evaporation from a reservoir in south-east England. Hydrological Processes 15, 2771e2778.
Flad, R., Zhu, J., Wang, C., Chen, P., von Falkenhausen, L., Sun, Z., Li, S.,
2005. Archaeological and chemical evidence for early salt production in
China. Proceedings of the National Academy of Sciences of the United
States of America 102, 12618e12622.
Geankoplis, C.J., 2003. Transport Processes and Separation Process Principles,
fourth ed,. Prentice Hall, New Jersey.
Glanville, J., 2005. Brine transport and woodland salt making in southwestern
Virginia: an hypothesis. Paper presented at the Eastern States Archeological
Federation Meeting, Williamsburg, Virginia, USA, November 12, 2005.
Joshi, V., Venkataraman, C., Ahuja, D.R., 1991. Thermal performance and
emission characteristics of biomass-burning heavy stoves with flues. Pacific and Asian Journal of Energy 1, 1e19.
Kanamori, H., Fujii, N., Mizutani, H., 1968. Thermal diffusivity measurement
of rock-forming minerals from 300 to 1100 K. Journal of Geophysical
Research 73 (2), 595e605.
Kostick, D.S., 2002. Salt. In: Metals and Minerals. U.S. Geological Survey
Minerals Yearbook, vol. 1. U.S. Government Printing Office, pp. 1e17
(Chapter 64).
Lange, N.A., 1952. Handbook of Chemistry, eighth ed. Handbook Publishers,
Sandusky, Ohio.
LeConte, J., 1862. How to Make Salt from Sea-Water? Governor and Council
of South Carolina, Columbia.
Lide, D.R. (Ed.), 1991. CRC Handbook of Chemistry and Physics, 72nd ed.
CRC Press, Boca Raton.

1462

D.G. Akridge / Journal of Archaeological Science 35 (2008) 1453e1462

MatWeb, 2007. Material Property Database. Available from: http://www.matweb.com (accessed 03.07.07.).
McCracken, J.P., Smith, K.R., 1998. Emissions and efficiency of improved
woodburning cookstoves in highland Guatemala. Environment International 24 (7), 739e747.
Olivier, L., Kovacik, J., 2006. The Briquetage de la Seille (Lorraine, France):
proto-industrial salt production in the European Iron Age. Antiquity 80,
558e566.
Penman, H.L., 1948. Natural evaporation from open water, bare soil, and
grass. Proceedings of the Royal Society, Series A 193, 120e145.
Penney, S., Shotter, D.C.A., 1996. An inscribed Roman salt-pan from Shavington, Cheshire. Britannia 27, 360e365.
Rife, D.L., Warner, T.T., Chen, F., Astling, E.G., 2002. Mechanisms for diurnal
boundary layer circulations in the Great Basin Desert. Monthly Weather
Review 130, 921e938.

Robie, R.A., Hemingway, B.S., Fisher, J.R., 1978. Thermodynamic properties


of minerals and related substances at 298.15 K and 1 bar (105 pascals)
pressure and at higher temperatures. In: USGS Survey Bulletin, 1452.
U.S. Government Printing Office, Washington.
Romanklw, L.A., Chou, I.-M., 1983. Densities of aqueous NaCl, KCl, MgCl2,
and CaCl2 binary solutions in the concentration range 0.5e6.1 m at 25, 30,
35, 40, and 45  C. Journal of Chemical and Engineering Data 28, 300e
305.
Sassaman, K.E., Rudolphi, W., 2001. Communities of practice in the early pottery traditions of the American Southeast. Journal of Anthropological Research 57, 407e425.
Shuttleworth, W.J., 1993. Evaporation. In: Maidment, D.R. (Ed.), Handbook of
Hydrology. McGraw-Hill, New York.
Williams, E., 2002. Salt production in the coastal area of Michoacan, Mexico.
Ancient Mesoamerica 13, 237e253.

You might also like