Sie sind auf Seite 1von 116

Sustainable Community Development Using Plasma Arc Gasification of Municipal Solid Waste

by
Celerick Stephens
A Thesis Submitted to the Graduate
Faculty of Rensselaer Polytechnic Institute
in Partial Fulfillment of the
Requirements for the degrees of
MASTER OF ENGINEERING SCIENCE AND
MASTERS OF SCIENCE IN MANAGEMENT

Approved:

_________________________________________
Dr. Ernesto Gutierrez-Miravete, Thesis Advisor
Rensselaer Polytechnic Institute
Hartford, Connecticut
April 2014

Copyright 2011, 2014


by
Celerick Stephens
All Rights Reserved

CONTENTS
CONTENTS ............................................................................................................................. i
LIST OF DEFINITIONS .......................................................................................................iv
LIST OF SYMBOLS .............................................................................................................vi
LIST OF TABLES ................................................................................................................vii
LIST OF FIGURES .............................................................................................................viii
ACKNOWLEDGMENT ....................................................................................................... xi
ABSTRACT .......................................................................................................................... xii
1. Introduction ....................................................................................................................... 1
1.1

Electrically sustainable communities .................................................................... 3

1.2

Waste sustainable communities ............................................................................. 4

1.3

Distributed Power ................................................................................................... 7

1.4

Matter reuse............................................................................................................. 7

1.5

Synergizing matter reuse and distributed power generation ................................ 9

2. Plasma Gasification Technology ................................................................................... 10


2.1

Plasma gasification ............................................................................................... 10


2.1.1 Plasma ....................................................................................................... 10
2.1.2 Gasification ............................................................................................... 11
2.1.3 Plasma gasification ................................................................................... 11

2.2

Plasma gasification for waste conversion ........................................................... 13

3. Methodology and Theory ............................................................................................... 15


3.1

Distributed scale plasma gasification .................................................................. 15


3.1.1 Municipal solid waste collection and disposal ....................................... 17
3.1.2 Tailored waste stream............................................................................... 18

3.2

Municipal solid waste gasification analysis ........................................................ 21


3.2.1 Community waste generation .................................................................. 22
3.2.2 Waste stream heating value ..................................................................... 27
i

3.2.3 Synthesis gas production.......................................................................... 29


3.3

Syngas conditioning.............................................................................................. 39

3.4

Energy production and storage ............................................................................ 41


3.4.1 Plasma arc system..................................................................................... 41
3.4.2 Fuel cell ..................................................................................................... 45
3.4.3 Waste feeding system............................................................................... 46
3.4.4 Electrical energy storage .......................................................................... 48

3.5

Ancillary systems.................................................................................................. 49

4. Results and Discussion ................................................................................................... 51


4.1

Plasma gasification model validation .................................................................. 51

4.2

Sustainable system sizing ..................................................................................... 58

5. Conclusions ..................................................................................................................... 61
5.1

Viability ................................................................................................................. 61

5.2

Notable errors and future work ............................................................................ 61

6. References ....................................................................................................................... 63
7. Appendices ...................................................................................................................... 68
7.1

Unsustainable middle class growth ..................................................................... 68


7.1.1 Middle class population rise .................................................................... 68
7.1.2 Middle class electrical power consumption ............................................ 70
7.1.3 Middle class waste production................................................................. 72

7.2

Proximate and Ultimate analysis of waste stream.............................................. 79

7.3

Plasma gasification program................................................................................ 80

7.4

Chemical properties .............................................................................................. 97


7.4.1 Heat Capacities ......................................................................................... 97
7.4.2 Gibbs free energy...................................................................................... 97
7.4.3 Heats of formation .................................................................................... 98

7.5

Single-stream recycling contamination ............................................................... 99


ii

7.6

Source reduction ................................................................................................. 100

7.7

Conveyor sizing .................................................................................................. 100

iii

LIST OF DEFINITIONS
Adsorption chemical process by which a chemical compound adheres to another
substance, allowing other compounds to pass without obstruction
Distributed power - generation of power on a where-needed and, often when-needed,
basis
Desulfidation process by which sulfur is removed from a substance (fluid)
Cathodic crucible a negative electrode shaped in a manner that allows for solid matter
to be contained within the electrode
Fuel Cell a device that coverts a fluid fuel to electricity by a chemical process which
separates hydrogen molecules, reforming it with oxygen and producing also heat
and water
Gasification - a process by which a hydrocarbon-based substance undergoes a thermochemical conversion from a solid to a gas
Globalization - the fundamental changes in the spatial and temporal contours of social
existence, according to which the significance of space or territory undergoes
shifts in the face of a no less dramatic acceleration in the temporal structure of
crucial forms of human activity (1)
Landfill(ing) - gathering wastes in aggregate and allowing generally natural processing
to decompose the matter over time (landfilling is the process thereby)
Leachate typically hazardous liquid composed primarily of water and contaminated
salts, present due to the transmission of water through waste
Middle class a segment of the population determined on the basis of an income (between $6,000 US and $30,000 US) that, on the worldwide aggregate year 2007
purchasing-power parity, provides household disposable income (2)
Municipal solid waste (MSW) Refuse or discard from a local populace that is in nature
composed of consumable, non-biologically generated matter
Organization for Economic Cooperation and Development (OECD) - an international
group of 50 generally affluent nations with the mission to promote international
and local policies to improve the socioeconomic situation of the worlds population
Plasma - a state of matter comprised of a charged fluid which is characterized by having
nearly equal concentrations of electrons and positively charged ions
iv

Pressure-swing adsorption the practice of introducing an adsorbent material to attract


a certain compound (or species) with the pressurization of gases
Proximate analysis method of determining the components in the waste stream on the
basis of the waste constituents and a percentage basis
Purchasing Power Parity theory in economics that seeks to set an equivalent exchange
rate by adjusting for the purchasing power of the exchanged currencies. This exchange is typically among national currencies and is relatable to the law of one
price, which states that barring fair trade vehicles (e.g. tariffs), the goods sold in
any country would be the same cost of the identical goods sold in any other country (2)
Refuse Derived Fuel (RDF) processed stream of waste matter that is useful for the
production of energy
Sanitary landfill designed areas of open waste decomposition that by way of hydrological and geological technologies prevent the emission of hazardous materials into
the environment
Sustainable (sustainability) a process that continues or progresses at a rate that promotes the repletion of resources that are consumed or changed by the process;
this ensures the process, or others that depend upon its input or output, may co ntinue without threat of decline
Ultimate analysis determination of chemical constituents of each of the components in
the waste stream based upon the proximate analysis
Vitrification a process in which amorphous materials are heated to a molten state and
cooled rapidly forming a glass. In the sense of this res earch it is used as a process by which non-gassifiable solids are encapsulated in glass by the process
described.
Waste material remaining after the transformation or extraction of the most readily
available or most useful content or energy contained within that material

LIST OF SYMBOLS
In order of appearance
Q Higher heating value of refuse derived fuel, in kJ/kg
G Gibbs free energy, or the standard free energy change of a species , kJ/mol
R Ideal gas constant, in J/mol-K
T Temperature, in K
K0 Chemical equilibrium constant for the stabilized reaction
K Equilibrium constant as related by the concentrations of the compounds
H Standard enthalpy change, kJ/mol
S Standard entropy change, kJ/mol
Partial change in the quantity the preceding symbol modifies
Partial derivative
p Pressure in N/m2
cp Specific heat given constant pressure, in J/mol-K
A, B, C, D, and E Coefficients of regression, used primarily for tabular reference
Hf Standard heat of formation of a chemical compound, subscripted, in kJ/kmol
nx Coefficient of a solute in the global gasification equation, in kmol
P Power, in kW
PH Conveyor power required to progress material (load), in kW
PN Conveyor power required to drive the conveyor (rotational), in kW
PST Conveyor power required to drive the conveyor (static), in kW
IM Media mass flow rate of the conveyor, in tonnes/hr
L Conveyor length, in m
Conveyor coefficient of friction
D Conveyor diameter, in m

vi

LIST OF TABLES
Table 3-1: Proximate analysis of the typical OECD municipal solid waste stream,
showing also the typical moisture content of constituents (10). *The water weight
of the waste presented is calculated from the data presented and is not present in
the original source. ................................................................................................... 26
Table 3-2: Water-gas shift reaction as experimentally demonstrated by Byun, et al (42).
The calculated production ratio is used for hydrogen reformation in the
distributed gasification analysis presented here. .................................................... 40
Table 4-1: Average composition of synthesis gas from the multivariate plasma
gasification analysis. ................................................................................................ 57
Table 4-2: Summary of the distributed plasma arc gasification system sizing based upon
power requirements and consumption, showing positive net power production . 60
Table 7-1: The growth of the world middle class compared to the growth of the world
population: a compilation of data from the Goldman Sachs* (2) and the U.S.
Census Bureau** (53). ............................................................................................. 69
Table 7-2: Elemental contents of a typical municipal solid waste stream (26)Categories
highlighted are represented constituents of the waste stream in this analysis.
**Mixed textiles are not found in Tchobanoglous text, but is averaged here for
simplification of the waste stream analysis. ........................................................... 79
Table 7-3: Heat capacities of elements and compounds of gasification reaction model as
compiled from the Chemical Properties Handbook (22) ....................................... 97
Table 7-4: Standard Gibbs Free Energy of Molecular Compounds Presented in the
Plasma Gasification Reaction (22) .......................................................................... 97
Table 7-5: Standard Heats of formation of molecular compounds presented in the plasma
gasification reaction (22) ......................................................................................... 98

vii

LIST OF FIGURES
Figure 1-1: Growth of the worldwide middle class from 1960 to 2010, with projections
to 2030. The total world middle class population is read on the left axis, while
the percentage is read to the right. Sources: US Census Beaureau (3), Wilson (2).
..................................................................................................................................... 3
Figure 1-2: Worldwide middle class electrical energy consumption, with data from 1960
to 2008 and projections to 2030. Compilation of sources: Goldman Sachs (2),
U.S. EIA (4), The World Bank (6) ............................................................................ 4
Figure 1-3: Worldwide middle class waste generation per year showing, with a
conservative projection that refuse generation rates are constant that more than
2.0 billion metric tons [2.2 billion tons US] of waste may be generated by the
middle class people of the world. The methodology and rationale for this
projection is detailed in Appendix 7.1. ..................................................................... 5
Figure 1-4: Waste stream processing showing the reduction of landfill waste from the
point of generation. Source: U.S. EPA (7)............................................................... 6
Figure 3-1: Plasma gasification process showing waste consumption and power
production cycle. The waste is first community generated, and proceeds through
a recycling stream prior to being converted into a refuse derived fuel. The RDF
is then converted to heat and waste gas in the plasma gasifier. The waste gas is
finally converted into electrical energy to sustain the plasma process and
supplement power for the community. In this figure are symbols describing the
constituents of the waste stream and waste gas that will be discussed in detail in
Sections 3.2 and 3.3.................................................................................................. 16
Figure 3-2: The waste stream constituents of waste generated in the United States in
2006 as determined from Chang (25). The category listed as 3% is other wastes
that are undetermined in Changs analysis. ............................................................ 23
Figure 3-3: Trends in the constituents of the municipal solid waste stream of the United
States from 1960 to 2006 as given by Chang (25), but simplified as shown more
directly in Figure 3-4................................................................................................ 24
Figure 3-4: Waste Stream Breakdown of the United States in 2006 as simplified from
Chang (25). Organics contains food and yard wastes. The category Textiles
includes rubber.......................................................................................................... 25

viii

Figure 3-5: Hydrogen reformation process from syngas, showing the system inputs to
providing for the water-gas shift reactions and the pressure-swing adsorption
system........................................................................................................................ 42
Figure 3-6: Plasma arc gasification system as concepted for small scale gasification. The
process runs only until the RDF is completely consumed, allowing for small
amounts of waste to be processed discontiguously and efficiently....................... 44
Figure 3-7: Block diagram detailing the concept of mass -based waste induction for the
distributed gasification of municipal solid waste. .................................................. 47
Figure 3-8: Grid-based electrical power storage technologies depicted by discharge
potential relative to system capacity. This information is presented from the
United States Department of Energy Sandia National Laboratories (48). ............ 48
Figure 4-1: Trends of electrical consumption (left) and hydrogen production (right) for
varying molar air ratios ............................................................................................ 51
Figure 4-2: The trend of electrical consumption (normalized) and hydrogen production
(inverted) to show optimum molar air ratio for a given waste stream. ................. 52
Figure 4-3: Hydrogen content of the syngas from the plasma gasification furnace as
related to the gasification temperature for a large sample of randomly diverted
waste stream constituents......................................................................................... 54
Figure 4-4: Plasma arc power requirements as a function of gasification temperature for
a large sample of randomly diverted waste stream constituents ........................... 55
Figure 4-5: Trends of the concentration of various waste stream constituents on
hydrogen gas production for a large sample of randomly diverted waste stream
constituents ............................................................................................................... 56
Figure 4-6: The waste generation rate and the plasma arc power requirements to yield
2.6 grams/second of hydrogen in the resulting syngas when three unique waste
streams are subjected to plasma gasification.......................................................... 58
Figure 4-7: Feed rate versus power consumption (upper diagram) and hydrogen
generation rate following gasification cycle and water gas shift (lower diagram)
showing the linear relationship among the three paramaters. The waste stream
has a contamination content of glass and metals due to imperfect separation. The
hydrogen flow rate is that leaving the syngas reforming system. ......................... 59

ix

Figure 7-1: Excerpt from Goldman Sachs report on the expanding middle clas s showing
a pareto of the per capita income in 2007 and comparing it to the world
projection in 2050. The chart on the right shows a swell of middle class income,
particularly in the nations known as the BRICs.................................................... 69
Figure 7-2: World population growth showing the growth of the middle class. Data
sources: Goldman Sachs* (2)and the United States Census Bureau** (53)......... 70
Figure 7-3: World electrical energy consumption on the basis of the world population.
This chart shows correlation to Figure 6.2, where the increase in the world wide
middle class correlates to the growth of the world projections of energy
consumption.............................................................................................................. 71
Figure 7-4: The projected world electricity consumption against the projected middle
class population growth showing close linear correlation between the global
middle class and power consumption...................................................................... 71
Figure 7-5: IPCC unrecycled waste generation data showing minimum annual rates of
carbon storage in landfills from 1971 to 2002. (8) Units of the ordinate axis are
teragrams (Tg) of carbon.......................................................................................... 73
Figure 7-6: PCC graphical representation of post-consumer waste generation between
1971 and 2002 (8). The units of the scales are in teragrams [Tg] ......................... 74
Figure 7-7: OECD waste generation in kg/capita/year for the period between 1980 and
2005. As all OECD countries are represented in the legend, only several are
listed. (9) ................................................................................................................... 75
Figure 7-8: U.S. EPA data showing municipal solid waste generation of the United
States between 1960 and 2008 (in million tons). (12)............................................ 76
Figure 7-9: U.S. EPA data showing municipal solid waste recycling (in million tons).
(12) ............................................................................................................................ 76
Figure 7-10: United States waste generation as determined independently by the United
States EPA (11)and the OECD (9). ......................................................................... 77
Figure 7-11: Waste generation as a function of the population of the world wide middle
class. This figure is analytically generated based upon data and p rojections from
the U.S. EPA (12), the OECD (9), and Goldman Sachs (2).See also Figure 1-3. 78

ACKNOWLEDGMENT
It is appropriate that the day of my final adjustments to this paper is on Easter Sunday. It is termed resurrection Sunday by many disciples of Christ. I look forward to the
completion of this study, and the completion of my degree, to be a resurrection of my
professional career. To Jesus Christ, my savior and Lord, and my Father God who sent
him and the Holy Spirit who dwells in me, I give all praise.
My wife Kimberly, whom I love dearly, has encouraged me more than significantly
in this endeavor. Her patience, understanding, attention, and love towards me are
unmatched in this world. I am ecstatic to have her as my wife.
To Dr. Antonios Mountouris, who provided me the details of the model which is the
backbone of this thesis, I give great gratitude.
I also thank Dr. Ernesto Gutierrez-Miravete, who has undoubtedly poured hours into
critically reviewing my work. He has been an excellent professor, coach, and mentor in
the completion of my Masters degrees. I feel that what he exuded as his personal
responsibility towards my degree completion inspired me to put in my best effort and not
let him down.

xi

ABSTRACT
The growth of a worldwide middle class and the electrification of mobility are factors stressing the current standards in electrical power generation. In developed
countries, electrical power is produced in mass at central locations and delivered over
large distances to communities. Growing electrical demand, coupled with inherent grid
losses, will ultimately require additional points of generation, further promoting the
consumption of valuable resources in power plant construction and operation.
Further, as middle-class communities continue to grow, the amount of waste generated also increases, promoting the problems associated with waste disposal. The
proliferation of inexpensive consumer electronics also ensures that a significant percentage of the generated waste stream becomes inorganic and biologically harmful, and the
innocuous disposal and treatment of post-consumer waste is a growing concern.
The research presented in this thesis addresses both of these concerns by investigating a means of converting waste into energy. Though waste-to-energy conversion is a
mundane technology, the peculiar benefit being proposed by this work is that energy
may be produced by the reduction of the waste stream on a local basis to advance
sustainable communities. The research and analysis presented develops a case for using
plasma gasification of municipal solid wastea clean method of converting waste into
useful synthesis gasto generate distributed power for individual communities. The
analysis conducted uses the principles of thermodynamic and economic systems analysis
to determine the benefit and feasibility of a distributed power plasma arc municipal solid
waste gasification system.
This thesis incorporates developed and vetted chemical modeling to simulate the
gasification of waste using plasma and sizes a scaled system that is capable of decrea sing electrical demand of a small community, while eliminating and neutralizing waste.
The results show that a community of fewer than 300 homes can fully support the
gasification process while producing enough power to offset the electrical demand of 10
homes. The distributed production of energy on a community basis is feasible, and it
could significantly and simultaneously reduce electrical delivery losses and waste
equating to an overall net energy savings.
xii

1. Introduction
Sustainability is an area of research targeted at reducing the ecological impact of
processes by reducing consumption and producing valuable inputs to itself or other
processes. The concept is made ever more important due to the growth the world
population and the finite resources by which to support it. However there is a global
phenomenon outstripping the resource pressures caused by population growth: it is the
growth of the worldwide middle class . The worldwide middle class currently is a
minority, but a person of this minority group consumes resources and produces wastes
many times that of the average person of the worlds population. Globalization 1 is a
factor supporting that growth. Globalization allows for larger pool of human resources
that can provide goods and services to distant regions conveniently (with little loss in
time) and, due to economic disparity in the costs of those rendered goods and services,
attractively on a global scale. In turn, the compensation for those rendered goods and
services provide for previously impoverished people to afford not only wealth to meet
their needs, but also wealth to satisfy their desires. Also, as the definition of globalization implies, this growing middle class is technologically driven; meaning that a greater
and growing percentage of the population enjoys the benefits of technology, especially
electrification. This places upon the established methodologies of energy delivery and
production stresses previously unmatched. Scientists and engineers are attacking the
issue by pursuing renewable power generation, giving life to new technologies in solar,
wind, biofuel, and tidal energy technologies. Conservationists are also pushing for more
efficient means of utilizing energy by pursuing advances that reduce power consumption, while preserving the comforts that electrical consumption afford.
A major sector of the environmental movement that has the potential to significantly
influence both energy production and energy (and resource) conservation is not as
1

It is this definition offered by Scheuerman (1) that demonstrates globalization and the
significance it has upon the world middle class population growth. This definition
highlights that the globalization is directly fueled by technology, which decreases the
significance of distance by decreasing the time by which physical and/or intellectual
goods or services subtend that distance.
1

popularly discussed: waste generation. Along with the growth of the world middle class
is an increase in world resource consumption. The waste generated by this increased
consumption middle class is also increasing: the space in which to dispose of the waste
is not. The generated waste must be disposed of, and that process requires the expend iture of energythe waste is generally transported to a destination and repeatedly
agitated or treated to accelerate its decomposition, as evidenced in landfilling operations.
Reductions in the energy required to dispose of waste is conventionally realized through
recycling, but this constitutes a fraction of the overall waste stream. The remainder (the
majority) of the waste is deposited in a landfill, with a small percentage used to generate
heating and electrical power. Because of environmental concerns however, waste-toenergy conversion is used primarily where land is a constraint.
The global middle class is also stressing the systems that deliver electrical power.
Even with the most advanced utility systems, significant losses are incurred in the
transmission of electrical power from the point of generation to the point of use. This
loss of electrical power through resistance and power transformation is compounded
with the distance over which that electrical power must travel. Again, referring to
Scheuermans definition of globalization, it is easy to see that the speed at which tec hnology enables the flow of ideas and materials effectively promotes the dispersal of
people, because the time-based perception of distance is decreased. The points of
electrical consumption are thus more widely distributed. The energy lost in transmission
is thus increased.
This thesis addresses waste disposal and its associated resource consumption, electrical power generation, and electrical power transmission with one unifying technology.
The unifying technology utilizes the process of plasma gasification to convert municipal
solid waste into energy. This technology is ecologically responsible in that it reduces
concentrations of both immediately hazardous and potentially hazardous waste products,
and is also sustainable as it requires a renewable resource, waste, as its fuel. The
purpose of this work is to explore plasma conversion technology as a means of reducing
the human waste stream and producing distributed electrical power to promote sustainable communities in an ecologically benign manner.
2

1.1 Electrically sustainable communities


The need for sustainable communities is readily apparent: though the rate of
growth of the world population is projected to be in decline for the coming decades , the
rate of growth of the middle class shows a progressive trend. Compared with only 16%
of todays world enjoying prosperity levels to the extent that some income is disposable,
in the year 2030 it is projected that 43% of the worlds population may be deemed to be
within the middle class 2 (this trend is depicted in Figure 1-1). This equates to 3.6 billion

Figure 1-1: Growth of the worldwide middle class from 1960 to 2010, with projections to 2030. The
total world middle class population is read on the left axis, while the percentage is read to the right.
Sources: US Census Beaureau (3), Wilson (2).

people in the year 2030 consuming electrical power at levels in excess of projected
capacity 3 (4) (as shown in Figure 1-2). Such an electrification growth rate is nearly the

Middle class percentage statistics are compiled from worldwide population statistics
from the United States Census Bureau (3) and the Goldman Sachs Economics Research
Group (2), respectively
3
This analysis was conducted based upon worldwide energy consumption data (6)
(4)and the projections of the middle class growth rate (2).
3

quintupling, in less than twenty years, the current electrical power generation rate of the
United States (5).

1.2 Waste sustainable communities

Figure 1-2: Worldwide middle class electrical energy consumption, with data from 1960 to 2008 and
projections to 2030. Compilation of sources: Goldman Sachs (2), U.S. EIA (4), The World Bank (6)

The growth of the worldwide middle class also influences waste production. In
2007, Americans alone produced 230 million metric tons [254 million tons US] of trash
at a rate of about 2.0 kg [4.5 lbs] per individual per day (7). The effect of this is noted
by research by the Intergovernmental Panel on Climate Change (IPCC) which directly
relates the rate of municipal solid waste generation to prosperity and population (8).
Assuming that the growing worldwide middle class may be represented by the consortium of primarily majority middle class nations known as the Organization for Economic
Cooperation and Development (OECD), the worlds middle class may generate municipal waste at a rate of 1.5 kg [3.4 lbs] per day (9). Though this waste generation rate is
substantially lower than that of the United States, projecting this over the growth of the
worlds middle class indicates that 2.0 billion metric tons [2.2 billion tons US] of refuse
may be discarded per annum in the year 2030, as shown in Figure 1-3. The amount of
4

waste generated in one year in 2030, based upon this potentially conservative projection,
would be enough to bury New York City in nearly 6 meters [20 feet] of trash 4.
Governmentally imposed regulation, and a shift in awareness in OECD countries ,
has led to the significant reduction in disposed waste because of waste source reduction,
recycling efforts, and conventional waste to energyincinerationtechnologies. Still
by the process described in Figure 1-4, 54% of the waste generated in the predominantly

Figure 1-3: Worldwide middle class waste generation per year showing, with a conservative projection that refuse generation rates are constant that more than 2.0 billion metric tons [2.2 billion tons
US] of waste may be generated by the middle class people of the world. The methodology and
rationale for this projection is detailed in Appendix 7.1.

middle class OECD nations is deposited in open waste decomposition sites called
landfills (7). These decaying landfills cause further environmental damage due to the
release of carbon dioxide, methane, and ground water pollution. Because of the dangers
of open waste decomposition, many nations have implemented regulations to allow for
sanitary landfills only. In these designed areas of open waste decomposition, hazardous
materials emission is limited by the implementation of hyd rological and geological
technology (10).

See Section 7.1.3


5

Figure 1-4: Waste stream processing showing the reduction of landfill waste from the point o f
generation. Source: U.S. EPA (7)

The emissions from landfills (both toxic and environmental) are significant. However, the process of disposing waste in a landfill is arguably more damaging. Municipal
waste vehicles travel to more than 79 million American homes, a 40,200 km [25,000mile] journey each year for the average was te collection vehicle. With 136,000 waste
collection vehicles in operation in the United States alone, the consumption on a fuel
basis exceeds 4.5 billion liters [1.2 billion gallons] of fuel each year (11)5.
The solution for ensuring sufficient resources exist to support the world demands is
multifaceted. Conservation, recycling, and the use of renewable resources for power
generation (wind, solar, and wave energy) are all important improvements in our energy
dependent society, but recycling is the predominate technology assisting in waste
reduction. The definition of sustainability should be augmented to not only address
power consumption and conservation but also waste generation and waste reduction.
Establishing waste and energy sustainable communities could significantly reduce
harmful emissions and improve resource conservation.

Municipal waste collection vehicles consume on average one liter of fuel for every 1.3
km [one gallon of fuel for every 3 miles] traveled due to the repetitive acceleration cycle
on waste collection routes (21)
6

1.3 Distributed Power


The majority of the power generated in the developed world is pumped onto an electrical grid, which is a means of distributing power over large expanses. The benefits of
this technology are that power can be generated where the resources are present, economical, and convenient. The detriment is that there is a significant amount of power
loss associated with large power distribution grids.
Line losses are a primary symptom of grid-distributed power. Seven-percent of the
electrical power generated in the United States is expended prior to consumption solely
due to transmission and distribution inefficiencies (12). This equates to 261 million
kilowatt-hours of electrical power, or enough to sustain 20,000 American homes with
electrical power per year6 (12). With growing demand and improved efficiencies, the
transmission and distribution loss fraction is not expected to increase significantly.
However a costly and resource intensive expansion of the electrical grid will be necessary to support the growth in electrical demand.
A better solution may exist, the distributed production of power. Distributed power
production is the generation of power on a where needed (and often times, when needed)
basis. Distributed power production reduces line losses merely due to the reduced need
for power transformation for long-distance transmission, as the power is generated near
the source of consumption. For this to be feasible, a local source of fuel and a local
means of converting it to electrical power is necessary. It would be of prime benefit if
that local fuel source were sustainably generated and transformed into electrical energy
with low ecological impact.

1.4 Matter reuse


All living organisms consume useful matter and expel wastes. Both useful
matter and wastes are in quotations as it is apparent in nature that the by-products
(wastes) of a given organism promote the continued life of another organism as a source
of fuel or habitation (useful matter). With the technologically advanced human organ6

In 2005, the average American single family detached home consumes electricity at the
rate of 13,162 kW-hrs. (4).
7

ism, however, more byproduct is produced than what is readily recyclable by natural
means. Because of this disproportionate cycle, humans are faced with the interesting
quandary of how to dispose of waste.
Though recycling and waste to energy methods have been used to reduce the waste
footprint of the developed world, the dominant means of waste disposal is landfilling.
Simply, the process of landfilling gathers wastes in aggregate and allows generally
natural processing to decompose the matter over time. Landfills in actuality are much
more complicated, and require a significant energy input. Firstly, the landfill must be
created. With environmental regulations to control the flow rate of biologically damaging leachates from landfills, wastes can no longer be openly discarded. Instead, landfills
are ecologically and geologically designed and constructed in a process known as
sanitary landfilling. Once the landfill is operational, the waste must be collected from
the points of generation and delivered to the site of disposal. To speed the process of
decomposition, increasing the useful life of the landfill, waste agitation is put into effect.
This process of using earthmovers to churn the wastes periodically results in yet more
fuel consumption. Upon attaining the limit capacity of a sanitary landfill, the landfill is
capped, a process by which organic materials (soils, grasses, other forms of vegetation,
and man-made inorganic materials) are placed on top of the landfill to render the site less
harmful.
Recycling in its various forms has a lower ecological impact and because of this fact
has been promoted worldwide. Today, up to thirty-three percent of the wastes that
would be bound for a sanitary landfill are recovered or immediately reused as otherwise
useful materials (11). Typically for the products that are in high demand, the reclamation of these materials is even more economical and environmentally friendly than
extracting the materials from the earth. Increased emphasis is being placed upon composting as well. In composting, organic materials are combined in such a way that the
natural decomposition of the materials is relatively fast. The byproducts are also nutrient rich, allowing for an environmentally friendly soil generation method. Though
recycling and composting currently make up a significant percentage of the recoverable
waste, vigilance and education are still required to ensure that the appropriate materials
are placed in the recycling and composting waste streams.
8

1.5 Synergizing matter reuse and distributed power


generation
Efforts have been made to utilize wastes efficiently. Incineration and landfill gas
recovery are means of waste reduction while generating electricity as a by-product.
Both cases however have undesirable results. In incineration, the wastes (after significant processing) are simply combusted as fuel. The byproducts of combustion however
are environmentally unsafe. Combustion methods have been found to release dioxins
and other potentially carcinogenic compounds into the atmosphere. Moreover, the
release of greenhouse gases (carbon dioxides and other gases) is undesirable. For the
collection of natural gas (methane) from decomposing matter, land (a precious resource)
must be occupied in great measure, and as previously discussed significant controls are
required to preclude the release of potentially hazardous materials into the ecosystem.
Posited in this research is that a synergism in methods of power generation and
waste stream reduction may exist. Imagine if the wastes generated in a locale were
directly used to generate at least part of the power requirements of that locale, with the
resulting byproducts being heat and innocuous building materials. Each of the processes
required for this synergism currently exists.
Plasma arc gasification, a process that has been proven in industry , is a means of
converting non-toxic and toxic waste streams into heat (useful for community and
process heating), process gases (which can be used for power generation), and in ert
building materials. This research further speculates that there is an optimum combination of plasma arc waste gasification and distributed power generation. In this solution,
the wastes that are locally generated fuel the process that generates local power. More over, this research posits that a more macroscopic benefit is also at play. Infrastructure
reduction, fuel and energy consumption, and electrical wastes are all significantly
reduced on a more global basis if more power were generated locally. Small-scale
plasma arc gasification of municipal solid wastes could be a key enabler of more sustainable communities in the developed world. The benefits may be yet more far
reaching in the undeveloped worlds, enabling electricity production on an efficient basis,
where currently no or limited energy resources exist.
9

2. Plasma Gasification Technology


2.1 Plasma gasification
Plasma gasification is a process of converting typically organic materials into ene rgetically useful substances, such as liquid or gaseous fuels. Though the industrial use of
plasma in material conversion is a relatively young technology, the discovery of plasma
has its roots in electric lighting. In 1927 an American chemist, Irving Langmuir, found
that an electrified fluid can transport electrons and ions in a similar manner as blood
plasma carries red and white blood cells. He stumbled upo n this discovery while attempting to prolong the life of the tungsten filament in a light bulb, and from this
discovery he was able to leverage his theories of this ionized gas to not only improve the
light bulb, but also to introduce a new branch of physics (13). It is upon the discovery of
plasma physics in which the root of this research is planted.

2.1.1 Plasma
The generally accepted scientific definition of plasma is a charged fluid, which is
characterized by having nearly equal concentrations of electrons and positively charged
ions. The physical characteristics of this matter is generally complex and not readily
described by the solid, liquid, or gaseous states of matterplasmas have also been
termed the fourth state of matter for this reason. Although the discovery of plasma is
recent, it is considered the most abundant state of matter in the universe, as it is the very
material composing our sun and countless other stars and celestial bodies in the universe.
On earth, we typically see plasma in the mundane forms of lightning and open flames,
and confined within fluorescent bulbs.
High-temperature plasmas that are formed by gases in electric arcs have also found
industrial uses. By passing high velocity gases in a highly charged electric arc, streams
of high-energy electrons and ions emit intense amounts of heat at temperatures exceeding 6,300 K [11,000F] (14). Devices that perform this function are plasma torches, and
at the exit conditions of a plasma torch, materials undergo rapid pyrolysis in the case of
organic materials or, in the case of inorganic materials, rapid melting , vaporization, and
10

even decomposition. It is this useful function of high temperature plasma that is the
basis of the technology described in this body of work.

2.1.2 Gasification
The other side of the technology presented is gasification. Gasification is a process
by which a hydrocarbon-based substance undergoes a thermochemical conversion from
a solid to a gas. This conversion allows for energy in a typically solid form to be tran sformed into a more useful form (either for transport, energy density, or utility).
Described as an ancient art by Higman and Van der Burgt due to its roo ts in woodfueled fires (15), the gasification typically described in engineering processes employs
pyrolysis, the heated and near-anaerobic conversion of a hydrocarbon to a useful,
combustible gas. The more modern development of gasification was employed in the
early 1800s as a means of turning coal into a form of gas. This town gas was widely
distributed for various uses ranging from industrial power and heat generation, to public
and private lighting, heating and cooking, in much the same way that natu ral gas is
currently delivered and utilized.
Because of the town gas origins of gasification and because of the widespread availability of the fuel source, coal gasification continues to be the most widespread
application of the technology. However, the quest for clean and renewable sources of
fuel spurred along with the discovery of liquid fuel processing using the Fischer-Tropsch
method of converting gas into liquid fuel (easing its transport and increasing its energydensity) has led to the gasification processing of other hydrocarbon feed stocks such as
biomass, including human wastes.

2.1.3 Plasma gasification


It is upon the convergence of the two technologies, plasma and gasification, that this
work is based. Generally, the process described here is hydrocarbon gasification. The
hydrocarbon is provided as the fuel source for the reaction and is derived from a solid
waste stream. The process of gasification is enhanced using high-energy plasma gas as
the pyrolysis catalyst. Both technologies, as described earlier, are proven in theory and
in practical application. Also the combination of the two processes (using plasma
11

torches to gasify refuse) has recently made the transition from the laboratory and controlled environments to industry implementation. The former and the latter reveal that
plasma thermal processing of human refuse is sufficient for the gasification of waste into
a useful, energy rich gas (synthesis gas) with about one-half the heating value of methane.
The combination of plasma and gasification differs significantly from waste incineration. Incineration processes are low-temperature thermal processes. Though the flue
gases generated by combustion often have sufficient free carbon (soot), hydrocarbons,
and carbon oxides (particularly carbon monoxide) to be combustible, the main product
of interest from incineration is heat. It is the heat from the combustion process that
provides steam to generate electrical power through steam turbine generators and
process or municipal heating. Further, incineration is typically an environmentally toxic
process. As wastes are destroyed at low temperature and in typically atmospheric
conditions, fly-ash, containing harmful substances, often escapes the process due to their
buoyancy and difficulty of processing. Further, the low-temperature recombination of
products of partial combustion results in the production of polychlorinated dibenzo-pdioxins, human carcinogens (16), and colloquially termed dioxins. To limit the production of toxic byproducts, incineration facilities employ extensive waste filtering and
recycling to remove potentially toxic products from the incoming waste stream and
extensive environmental monitoring and flue gas and ash processing to remove harmful
impurities from the exhaust.
Conversely, plasma gasification is regarded in industry as a clean process for the
thermal conversion of waste to fuel. Because of the extremely high temperatures
involved and the oxygen starved environment under which the process takes place,
municipal solid wastes can be nearly instantly converted into a combustible and useful
fuel with significantly fewer potentially harmful agents being generated (17). Further,
conventional gas scrubbing technology may be used to process clean the synthesis gas
while reprocessing the syngas contaminants through the plasma gasifier. The resulting
synthesis gas with a heating value about half that of natural gas is immediately useful for
the clean production of steam for steam power generation, directly in a Brayton cycle, or
as a fuel source in a fuel cell. There is only one other form of waste product from
12

plasma gasification. The inorganic materials, which would generally become fly ash in
an incineration process, become molten materials due to the high process temperature.
These molten materials are vitrified, encapsulating even potentially harmful substances
in an inert silica or alumina glass, while other hazardous materials at these process
temperatures form harmless oxides, which are also vitrified (18). This innocuous glasslike material is useful in construction materials, without any subsequent treatment.

2.2 Plasma gasification for waste conversion


Plasma gasification is a technology that can be applied to a range of input fuels. In
fact, because byproducts of plasma gasification have low harmful substance percentages
and potentially hazardous solids are vitrified, plasma conversion is regarded as a means
of transforming hazardous wastes (primarily biological and nuclear wastes) into more
benign forms. In these processes, the goal is the reduction of waste and the benumbing
of any hazardous byproducts. Though one of the outputs of the conversion is synthesis
gas, the high concentrations of inorganic matter in these waste streams prohibit the
production of significant quantities of gas for effective energy production. However
hazardous waste is minimized and nearly eliminated, therefore the overall cost associated with the application of plasma technology to the conversion of this hazardous waste is
justifiable (19).
Coal conversion using plasma was among the first explored uses of plasma gasification for the production of energy. It provides a clean process gas for electricity
production while reducing dioxin and carbon emissions. The plasma conversion of coal
also vitrifies the trace contaminants in coal (sulfur, metals, stone and sand) into a useful
construction material. It comes as well with the benefit of facilitating the retrofit of
existing coal-fired power plants while reducing upgrade costs (20). Coal gasification has
a significant side effect: the coal that is to be used in the process needs to be delivered
from its place of origin to the point of power production. Moreover, at the place of
origin, large amounts of energy are expended for the extraction of coal. With an overall
consideration, the use of coal for the production of energy using plasma gasification
involves a net energy loss.
13

The use of plasma gasification presented here may be considered of three purposes.
The primary area of focus is sustainable power production. The secondary focus also
builds upon the premises of the first, the elimination of waste at the source of the waste.
Finally, by generating the power at the place of necessity, significant reductions in
power transmission can be realized, and by consuming waste locally to generate power
locally, significant reductions in waste energy expenditures are also attainable.

14

3. Methodology and Theory


3.1 Distributed scale plasma gasification
The process of plasma conversion of municipal solid wastes (detailed in Figure 3-1)
starts with the community generated waste stream. The waste stream is composed of
post-consumer and post-industrial materials with perceived low residual value. In the
common culture of the developed world, the low value waste is easily discarded, either
by separating the contents of the waste stream to remove those components of higher
residual value by recycling, and either incinerating or landfilling the majority of what is
remaining. In both casesrecycling and discardingthe waste is centrally gathered,
necessitated by economics, legal and environmental controls, and public desire 7. The
plasma gasification process envisioned here begins at the point of generation. Instead of
gathering waste and transporting it to a central processing facility, the waste is transported only a short distance to a distributed collection facility, which processes then only a
fraction of the waste that would be generated by a municipality.
The second step in the gasification process is waste stream conditioning. The waste
stream is filtered for content that ultimately detracts from the desired result, the produ ction of synthesis gas and heat. The need for this step in the process resulted from a
macroscopic optimization showing the desirable products of the process (electricity and
heat generation and waste reduction) significantly benefit from conditioning the waste
stream. Recycling materials and removing moisture from the waste stream are both
aspects of waste stream conditioning. The process presented attempts to maximize
electrical output by balancing the moisture content and constituents of the waste stream.
This processed waste stream then becomes fuel (known in industry as an RDF,refusederived fuel) for the plasma convertor.

Waste collection is a highly regulated industry. Sanitary and environmental concerns


have led to legislation that specifically zones where and how waste will be stored. This
is also driven by public desire to not have waste collection facilities in close proximity to
developed areas. The importance of these factors will become apparent as they may be
detractors to the adoption of a distributed power generation scheme similar to that
described in this research.
15

Figure 3-1: Plasma gasification process showing waste consumption and power production
cycle. The waste is first community generated, and proceeds through a recycling stream prior
to being converted into a refuse derived fuel. The RDF is then converted to heat and waste gas
in the plasma gasifier. The waste gas is finally converted into electrical energy to sustain the
plasma process and supplement power for the community. In this figure are symbols describing the constituents of the waste stream and waste gas that will be discussed in detail in
Sections 3.2 and 3.3

The refuse-derived fuel is then passed into the plasma convertor. In the plasma
convertor, the waste fuel is thermally converted at extremely high temperatures resulting
in a gas which, when controlled in its cooling, produces a combustible fuel. Another
byproduct of the plasma converter is a molten slag which is process cooled in a controlled manner to ensure vitrification, the result of which is an inert solid that needs no
further processing. Finally, due to the heat input into the waste stream by the plasma
torch and the energy rejection from the synthesis gas and vitrification processes, a
significant amount of residual heat is generated. This byproduct can be fluid transferred
16

for use in waste conditioning (reducing moisture content) and community heating (for
heated water, process heat, or for air-conditioned living spaces). It is the primary
byproduct, synthesis gas, which is of significant interest as it becomes the main source
of energy for the plasma process.
The synthesis gas, or syngas, can be combusted or directed to a fuel cell (as is done
in this analysis) for the generation of electrical power. The electrical power created from
the syngas is at least sufficient for creating the plasma arc, but it is also possible that
enough syngas is produced to power the equipment used to create the RDF and treat the
syngas while providing supplemental power for the community. In this way, not only is
the ultimate amount of refuse reduced, but power is also generated to realize the benefits
of waste reduction and distributed power production.

3.1.1 Municipal solid waste collection and disposal


The conventional collection and transportation of municipal refuse is a cycle that itself generates significant waste. In most municipalities in the United States and other
developed nations, refuse vehicles transit neighborhoods and businesses to collect wastes
at the point of generation and deliver them to a more centralized disposal location. As
noted in 1.2, the refuse collection cycle is energy intensive. The United States consumes
enough fuel in waste transit vehicles each year to power 2.2 million cars during their
daily commute 8. Though there are recent efforts worldwide to address the environmental impact of waste collection vehicles, they are primarily aimed towards hybrid
technologies and alternative fuels such as natural gas and hydrogen combustion vehicles.
Recently in China, even battery electric vehicles have been introduced to address the
need to reduce fuel consumption. Efforts to reduce fuel consumption in refuse vehicles
(and in general large vehicular diesel engines) are being regulated by legislation. In the
United States in 2010, heavy-duty engine emissions standards were implemented specifically to address this concern. It is estimated that if 50% of waste collection vehicles in

The calculations are based upon the American average fuel economy of 0.14 L/km
[17.2 mpg] and the American average yearly commuter vehicle mileage of 19,000 km
[11,900 miles]. (58)
17

the United States were converted to natural gas, the savings could amount to an oil usage
reduction of 14.3 million barrels of oil each year (21).
Regardless of the potential savings in reducing the oil consumption of refuse veh icles, the collection and transportation of waste perpetuates the waste of resources for the
sole purpose of disposing of waste. Another solution to the problem is to diminish the
need for the collection and transportation of waste. By consuming the waste locally, at
or very near the point of generation, significant reductions in overall fuel consumption
may be realized. For a typical waste stream in the United States, more than 58% of the
waste by mass9 is plastic, paper, and organic non-food wastes (wood, leather, textiles,
and rubber). Local plasma gasification could remove these wastes from the waste
stream, resulting in a highly recyclable mix of glass and metals. Accordingly, the energy
consumption associated with waste transportation and collection of the remainder of the
waste stream is more than halved. This opportunity is addressed in the distributed power
aspect of plasma waste gasification explored in this research, as most of the waste
generated could be locally consumed in a small, local plasma gasification facility.

3.1.2 Tailored waste stream


Detractors of plasma waste gasification argue that the technology is detrimental in
that it promotes the waste generation mentality while also degrading recycling efforts.
The argument is posed that if all waste is recoverable there is no benefit in waste reduction, because waste is a source of fuel in the new paradigm. Also the advantages of
recycling are obsoleted because paper and plastics wastes are even better fuels for
plasma gasification. There is truth in these claims, and a thorough review of the application of distributed plasma gasification also promotes recy cling, as well as waste
reduction 10.

A mass basis is considered due to the assumption that compaction reduces the volume
of trash collected.
10
Source reduction as it pertains to waste-to-energy methods such as plasma gasification
is addressed in the Appendix in Section 7.6.
18

This research shows that a tailored waste stream creates a more energy rich synth esis gas. Elementarily, one can come to this conclusion by considering the constituents of
a gas with a high recoverable energy. On the basis of the higher heating value of a fuel,
hydrogen is the most energy rich, with a heating value of 142,000 kJ/kg [61,100 BTU/lb]
(22). A more abundant natural gas, methane, is a hydrocarbon with a chemical compos ition of CH4 and has a heating value of 55,600 [23,900BTU/lb] (22). When organically
derived components of the waste stream are subjected to plasma conversion the result is
the generation of hydrogen and low heating value hydrocarbon gases, while the inorganic portions of the waste stream only debits the net energy generation of the process. In
fact, the most energy productive plasma waste processing would completely recycle out
all metals and non-organic wastes in favor of dry hydrocarbonssuch as paper, plastics,
rubber, textiles, and wood. It should be noted in the previous statement that dry hydrocarbons are preferred: the process of converting water into its gaseous form in plasma
gasification only consumes energy with no significant benefitwater effectively only
changes its form without releasing valuable byproducts by the plasma process. For th is
reason wet organic wastes such as food wastes only subtract from the net valuable
energy of plasma gasification of waste. Since one of the goals of plasma gasification is
the production of the highest energy synthesis gas for the production of electricity,
recycling on a plasma waste conversion basis simplifies the waste stream into the simple
categories of food waste, metals , glass, and everything else.
It is this simplification of the recyclable waste stream that continues to promote
plasma gasification. To reduce the costs of recycling collection, and to promote recycling collection from the private and corporate sectors, many municipalities have
transitioned to single-stream recycling. In single-stream recycling, all post-consumer
content is aggregated at the time of collection. The public remains uneducated about
what is acceptable as recyclable materials, and this method has resulted in more contaminated recycling streams, as inappropriately separated materials degrade the quality of
recycled input leading to material losses. For example, in multi-stream recycling there is
no differentiation in paper products, so all articles from copier paper to milk cartons are
bundled together at the single-stream recycling centers (or material recovery facilities)
and then sold to pulp and paper mills. Though this is in theory beneficial, large amounts
19

of what is thought to be recyclable is actually discarded at the pulp facilities due to


contamination as the single-stream papers include plastic and wax coated papers (e.g. as
from plastic view windows in postal envelopes, plastic coated newsprint, and modern
milk and juice cartons which are wax coated with the plastic pour spouts and caps).
While loss rates for metals and glass are low (7% to 12%) loss rates for plastics and
paper derived products are 20% to 40% (23). The contaminated wastes are ultimately
incinerated (conventionally) or land-filled. However, in the plasma gasification model,
those wastes with high reclamation rates (metals and glass) are still recycled, while the
wastes that are more easily contaminated (plastics 11 and paper) are plasma converted
into synthesis gas and the contaminants are inconsequential to the process. Using
plasma gasification for waste treatment allows for an improved recycling methodology
in this respect.
Another cost of recycling is also averted by converting waste to energy using pla sma gasification on a local basis. In most municipalities, recycling input is collected in
the same way as municipal wastelarge waste transfer vehicles are employed to
transport the waste to a central processing facility. Where single stream recycling
collection is employed, the waste is mechanically (and even manually) separated into the
various recycling streams at the expense of additional resource consumption. By reducing the recycling to glass and metals, only 30% of what is currently recycled or
composted would need to be transported to a central processing facility (24).
From these two counterarguments, we see then that recycling in waste management
is not degraded, but instead is newly defined 12. Plastics, regardless of contamination
rates, being hydrocarbons of significantly lower moisture content, boost synthesis gas
generation rates and raise syngas heating values. Even the scenario of plastic or paper
based containers retaining other products (from inadequate cleaning or emptying of
containers) is addressed in that the plasma temperatures are sufficiently high to gasify
11

Plastics recycling initiatives result in significant wastes due to contamination. Because the public remains unaware of the necessity to empty thoroughly the plastic
containers to be recycled, quantities of food wastes, oils, cleaning fluids and medications
pollute the potentially recyclable plastics stream.
12
Source reduction is an important consideration in waste management and is discussed
in the Appendix Section 7.6.
20

the container and the residual contents while vitrifying constituents that do not gasify. It
is possible that paper and plastic wastes with high moisture contents (as in a con tainer of
water being introduced in the plasma convertor) increase the energy required for gasification, but the only consequence in such a scenario is the net loss in power; no waste is
created. In fact, due to the ultimate compounds present, more useful hydrogen gas may
be produced. This also holds true for other recycling and non-recycling content that may
be introduced into the waste stream.
Metals and glass13, are considered undesirable as the power for melting and vaporization during the plasma gasification process produces no useful synthesis gas. Multistream recycling in this scenario is then reduced to two streams of low contamination
rates (or conversely high recovery rates). All other forms of waste provide for the
creation of a useful syngas of moderate heating value in the plasma gasification reaction.
The robust nature of this recycling scheme precludes public education and accidental
recycling stream contamination, and for this reason plasma waste conversion of post consumer waste products may be viewed as enhanced recyclinga recycling method
that generates the ultimate starting state of the matter being consumed: energy. For its
benefits to recycling, plasma gasification of waste might be considered plasma enhanced
recycling.

3.2 Municipal solid waste gasification analysis


The key differentiator of waste to energy in plasma gasification is the use of an energized gas to promote waste conversion. It is this process that heats the waste stream so
quickly and to such high temperatures that the waste is not burned. The plasma torch
leads to the decomposition of the primarily hydrocarbon waste without the combination
of oxygen that would promote combustion. Also, the temperatures are so high that
inorganic wastes are melted allowing for vitrification of potentially harmful substances.
Separating plasma pyrolysis from lower temperature pyrolysis is the significant redu cGlass and silicates typically comprise 5% of a municipalitys waste stream (24). In the
ideal case, all of this too would be recycled, however reality requires some small amount
of this to be present in the plasma processing because it is the substance that allows for
the vitrification of solids (both toxic and non -toxic).
21
13

tion in the formation of carbon, soot or char. The analysis of this process is of prime
importance in the plasma gasification of waste because it generates the synthesis gas that
makes this process ecologically and economically feasible.
The plasma furnace (also referred to as the plasma convertor, as it changes waste into useful byproducts) is the fundamental component in the high-temperature pyrolysis of
the waste stream. The process for determining the value of the plasma gasification
process is based upon a chemical ultimate analysis. The products of the ultimate analysis
are the gases that are used to then generate electrical power, which is in turn used to
power the plasma gasification system. The residual electrical power is then used to
supplement the power requirements of the community. The residual heat from the
plasma torch may be used to dry the input waste (reducing the moisture content and
increasing the net energy output) and can be used for hot water production for process or
community usage.

3.2.1 Community waste generation


The analysis begins with the input waste stream. It is important to understand the
amount of waste a community typically generates to determine the minimum size of the
community that can be supported with the plasma gasifier. In the United States, 2.0 kg
[4.5 lbs] of waste is generated per day per person excluding recycling efforts 14. This
generation rate is used as a global estimate of the middle class waste production rate and
intentionally does not include the reduction of waste due to recycling efforts as the paper
and plastic content is key in increasing the energy output of the gasification cycle. It is
important to understand the constituents of the waste stream and the chemical compos ition for determining the products of gasification.
The content of the average United States waste stream was quantified in research
presented by Shoou-Yuh Chang (25) and is shown in Figure 3-2. The trend of the
change in materials in the waste stream over time is significant, as shown in Figure 3-3.
Changs work shows that from 1960 to 2006, plastics have grown as a disproportionate
14

Section 7.1.3 details waste and recycling statistics used in the determination of the
waste and recycling habits of the worldwide middle class.
22

piece of the waste stream (due in large part to products, packaging, and electronics).
Paper has recently remained unchanged or is on a slight decline as a percentage of the
overall waste stream. The importance of this trend is that plastics and paper provide the
highest calorific benefit in the plasma conversion of waste. As a part of the optimization, it may also be determined that the presence of plastics provides for the higher
production of synthesis gas.
A simplification from Changs work is made in the use of the data presented , noted
in Figure 3-4. As, textiles and rubber and leather are of similar percentages, and noting
that the presence of metals, rubber, plastics, and inorganics are also present in todays

Figure 3-2: The waste stream constituents of waste generated in the United States in 2006 as
determined from Chang (25). The category listed as 3% is other wastes that are undetermined in
Changs analysis.

textiles, these categories are combined into a single category, textiles, in this analysis.
Similarly, due to the likeness of food wastes and yard wastes, these two waste streams
are combined into a single category termed organics. Thought was given towards
combining other similar categories (wood and paper in particular) but this was not done
due to the significant chemical differences that may exist due to the methods of production treatment and processing for wood wastes (e.g. paints and stains, waxes, and
plastics).
23

Figure 3-3: Trends in the constituents of the municipal solid waste stream of the United States from
1960 to 2006 as given by Chang (25), but simplified as shown more directly in Figure 3-4

Obviously missing from the constituent analysis (or proximate analysis) of the
waste stream is the presence of water. This is accounted for in the typical content of
moisture inherent in the constituents, as helped by the relationship developed by
Tchobanoglous (26) and summarized in Table 3-1. The content of moisture will be
important not only for accounting for the water present in the waste stream, but also for
determining the amount of drying that may help optimize the production of synthesis gas
and in the raising of the temperature of the plasma reaction. The basis for determining
water used in this analysis is simply the product of the constituent mass and the typical
moisture content. Research presented consistently implies that this method is an underestimate of the moisture content of the waste stream. It must be noted that in this
analysis, the presence of externally provided water in the waste stream (from the colle ction of rainwater and waste water) is expected to be minimized, as the waste will be
locally generated and disposed. Water that enters the waste stream from large scale
collection methods can therefore be discounted. As such, the starting input moisture of
the waste stream is nearly 20%, as opposed to the 28% 35% moisture content docu24

Figure 3-4: Waste Stream Breakdown of the United States in 2006 as simplified from Chang (25).
Organics contains food and yard wastes. The category Textiles includes rubber.

mented in the prevailing research. Further, the conditioning of the waste stream (prima rily in the composting or other disposal of organic wastes) will drop the moisture content
of the waste stream to approximately 7%.
This reduction of the moisture content of the waste stream at the point of generation
is a very significant point: the presence of water increases the input energy required to
maintain a given pyrolitic temperature. More energy may be required for electrical torch
operation and waste stream drying. There is a balance involved, however. Water is
required to aid in the formation of hydrogen, methane, and other low-energy gases that
comprise the syngas, which will ultimately be providing electrical power fo r the gasification reaction and for supplemental power generation. This study will determine this
balance.

25

Table 3-1: Proximate analysis of the typical OECD municipal solid waste stream, showing also the
typical moisture content of constituents (10). *The water weight of the waste presented is calculated
from the data presented and is not present in the original source.

Weight
Generated
Material
Paper & Paperboard
Glass
Metals
Plastics
Rubber & Leather
Textiles
Wood
Other Organic Wastes
Other Inorganic
Wastes

77.42
12.15
20.85
30.05
7.41
12.37
16.39
64.69

Percentage
Generation
31%
5%
8%
12%
3%
5%
7%
26%

Typical
Moisture
Content
6%
2%
0%
2%
15%
6%
35%
60%

8.28

3%

0%

(Millions of
Tons)

Water
Weight
of
Waste *
(Millions
of Tons)

4.65
0.24
0
0.60
1.11
0.74
5.74
38.81
0

Finally, because the optimization of power generation from the waste stream is d ependent upon the energy of the input fuel, we need a means of relating the contents of
the waste to the chemical constituents of the waste stream. A chemical elemental
analysis (ultimate analysis) of the waste stream generates this data. As it is beyond the
scope of this research to perform the ultimate analysis in a laboratory environment, again
the research documented by Tchobanoglous for the dry basis of elemental content of
items in the waste stream is used and compared with the works of Boie. Explanation of
this method is further developed in Section 3.2.2. Useful in this analysis is the
Tchobanoglous (26) break down on an elemental basis the predominant contents of a
municipal solid waste stream, shown in Appendix Section 7.2, Table 7-2.
This is further backed by an ultimate analysis that classifies the waste stream in
terms of its heating value, a measure of the potential calorific energy of the refuse
derived fuel. Measuring this requires the use of a bomb calorimeter in a controlled
laboratory environment. Instead, the conditioned waste stream that results from selective recycling will be compared to that presented in research, ensuring the analysis
remains valid over the range of conditions being evaluated. Primarily the Boie methodology is employed to provide this assurance.

26

3.2.2 Waste stream heating value


To determine the theoretical synthesis gas production, the method by A. Mountouris
(27) was originally employed. However, because Mountouris work focuses mainly
upon an analyzed waste stream (the heating value of the particular input waste stream
was empirically derived in a laboratory environment) another method of determining the
calorific value of a constituent-variable waste stream is required. Schuster, Lffler,
Weigl and Hofbauer (28) were presented with a similar problem in understanding
biomass steam gasification. The answer in Schuster, et al. research was the Boie relationship, which they found to have good correlation in high moisture environments .
There are several methods of estimating the heating value of organic fuels, but two
leading methodologies are frequently repeated in gasification research, the Boie equation
and the Dulong relationship. The Boie equation (29) is an empirical derivation developed in 1953 to estimate of the lower heating value of mixed wastes (30), but is more
typically applied for coal gasification. The Boie equation has been determined to be
accurate when applied within oxygenated (more than 10% oxygen) and moist (more than
10% water content) environments (31). It should be noted, however, that Dulong (32)
has been shown to correlate very wellwithin 3%when applied to refuse derived fuel
from municipal solid waste on a dry, ash-free basis. Interestingly, the Dulong equation
has also been shown to have significant error (31) when empirically correlated to oxygen
rich environments (where oxygen is not only present in the waste, but introduced from
the environment) and moist environments. When applied to refuse material, Rigo (30)
also found that the Dulong relationship is less accurate than Boie.
Siegel, et al. (33), vetted the Boie relationship in research presented by determining
the heating value of biomass using a bomb calorimeter. His work showed that Boies
method employed in an ultimate analysis produces sufficiently precise determination of
the lower heating value of biomass which is, on an average basis, 2.5% of the empirically determined upper heating value.
Given the diversity of results regarding the various heating value determination
techniques, the Boie equation is used in this analysis. The analysis varies the waste
stream dramatically to develop trends that help determine the best constituents in the
27

waste stream that promote hydrogen production while requiring low input energy for
gasification. The accounting for the presence of nitrogen15 (32), oxygen, and potentially
significant amounts of moisture makes the choice of Boie a robust method for this study.
The robustness relative to these constituents is important due to the design of the gasification furnace: the plasma arc operates in a natural air furnace to reduce cost and
complexity of operation.16 Without laboratory testing, the work of Boie allows for the
determination of the heating value of the waste stream, necessary for the equilibrium
reaction that results in the production of synthesis gas .
The accounting for dry ash is also important, as the ash in the process is also vitrified in the slag produced from the plasma process. In this analysis, ash is accounted for
in the proximate analysis of the waste stream as noted in the Appendix, Table 7-2.
The completion of the proximate analysis and the ultimate analysis, as described earlier, is the
starting point for employing the Boie relationship. The Boie formula for the higher-heating value of
the waste stream as a fuel is given by [1]

[1, where C, H, O, N and S are the elements, Carbon Hydrogen, Oxygen Nitrogen,
and Sulfur and the coefficients are the percentages of the constituents of the waste
stream from the ultimate analysis , and Q, the higher heating value of the RDF is given in
kJ/kg.

[1]

In the work presented by Mountouris (34), a bomb calorimeter was used to experimentally determine the heating value of sewage sludge. Several widely accepted
mathematical models were used to approximate the sludge in order to select an analytical
15

Buckley (32) indicates that the presence of nitrogen in the waste is small enough that
it may be discounted. However, the presence of nitrogen in air (for an atmospheric
process as is designed here) may be significant.
16
While an atmospheric environment of the plasma furnace promotes inexpensive
facilities and allows for the simplicity of operation, it is noted that the facility will have
the propensity to generate more ashan effect that can be countered by increasing the
plasma furnace temperatureand require more frequent maintenance. The useful life of
the plasma torch is also reduced due to oxygenated erosion of the electrode.
28

method that could be used to determine the HHV of the specific sludge of his analysis.
The Boie relationship was not referenced among them, but it yields results within 4
percent of the empirical value of the dry, ash-free basis of the Psittalia sludge waste
stream referenced 17. The analyzed sewage-sludge waste stream consisting of 55%
carbon, 8.0% hydrogen, 33% oxygen, 3.8% nitrogen, and negligible sulfur content on a
dry, ash-free basis, gave a higher-heating value of 24,198 kJ/kg, and the Boie relationship for the same waste stream provides a heating-value of 25,100 kJ/kg, an overestimate
consistent with that found in other research (31) (32). The difference in the two methods
is significant and will be discussed in the conclusions (Section 5.2) of this analysis.
To ensure that the Boie method is applicable to the municipal solid waste stream of
this analysis, the ultimate analysis of a recycling-optimized (or conditioned) waste
stream is compared to that of the sewage sludge. It will be shown later that the cond itioned waste stream desires to be free of dry paper (see Section 4.1), glass, and metals,
yielding a fuel that is 52% carbon, 6.7% hydrogen, 33% oxygen, 2.2% nitrogen, and
negligible sulfur content. Each of these elemental constituents is within 2 percent of the
sewage sludge on a dry basis, providing confidence in the application of the Boie
methodology to the determination of the heating value in this analysis.

3.2.3 Synthesis gas production


The methods of thermochemistry18 are used to analyze the production of the synthesis gas from a global gasification reaction of a single-chambered furnace.19 From the
prior sections, the conditioned waste stream (or refuse derived fuel) is determined on an
elemental basis, determining the amount of carbon, hydrogen, oxygen, and nitrogen
present in the waste. The content of sulfur is disregarded due to the small amount of
sulfur present in the reaction. In a later section, however, it will be shown that the
17

Mountouris, et al (34) shows the Dulong relationship to have 0.6% error from the
calorimeter results, on a dry, ash-free basis.
18
Credit for the insight and application of the theory discussed in this section is due to
A. Mountouris et al. (27), as the methodology is concisely demonstrated in their study of
gasification and the development of the GasifEq model that is the basis of this analysis.
19
Other multi-phased approaches to plasma gasification (35) are being developed that
will yield higher power outputs by employing separate gasification chambers, but this
also introduces other system complexity and cost.
29

presence of even a small amount of sour-gas cannot be tolerated by the fuel cell. For
this reason, the syngas filtration system includes energy debits for having to remove
H2S, along with CO and CO2 for the longevity of the fuel cell stack.
The plasma gasification process is also made to occur in environmental conditions.
The primary benefit of this is the absence of the need for a controlled atmosphere
(temperature, pressure, and inert gas supply) which significantly reduces the cost and
complexity of the plasma reactor unit. The major detriment is the significantly reduced
life of the plasma torch or electrode. The chemical reaction is established to make use of
the free oxygen and nitrogen in the atmosphere.
The moisture in the waste stream that typically occurs due to distributed waste collection and central gathering, also allowing for the collection of atmospheric
precipitation, is disregarded. Instead, moisture in the waste stream is a result of only the
constituents of the waste stream and is calculated from the conditioned (post-recycled)
trash stream. The presence of atmospheric moisturerelative humidityis also not
accounted for in the moisture content of the waste. The atmospheric temperature is
accounted for in the analysis and is factored into the ambient temperature of the reactor
as an average yearly temperature for the New England region of the United States
(10.7 C).
The gasification reaction proceeds from the proximate analysis (the content of the
waste stream, including moisture and ash content) to the ultimate analysis (the elemental
constituents of the waste stream) and the determination of the heating value of the waste
stream (which is, with the pre-gasification recycling and waste shredding, the refuse
derived fuel) to the thermochemical equilibrium equation. The method of performing
this analysis is embedded in the GasifEq model developed by Mountouris et al. (27)
This method employs the ultimate analysis of the waste stream (

), where ash,

sulfur, metals, and the other trace elements, if present after the gasification reaction, are
assumed to be mostly vitrified in the conversion of waste material into p rimarily gaseous
forms. However, the presence of these trace compounds is important for the environ-

30

mental performance of the reaction and the products of significance are simplistically 20
handled in this study as a treatment of the syngas prior to induction in the fuel cell (as
detailed in Section 3.3). The global gasification reaction is:

[2]

In this equation, the coefficients n x denote the stoichiometric concentrations in kilomoles (kmol) to balance the equation, m is the amount of air per kmol of waste and w is
the moisture content, also in kmol. The amount of air is an important factor, as it
directly controls the oxidant. Another significant oxidant comes from the waste stream
itself. Mountouris et al. (27) noted that the increasing content of oxygen significantly
decreases the production of hydrogen (as more free oxygen is present to promote the
formation of water) (35). Alternately, increasing oxygen lowers the required input
energy, as anticipated due to the oxidant nature of the reaction. For the reasons stated,
the molar concentration of air governing the stoichiometric ratio will be fixed 21 as to
produce the most hydrogen for the least consumed energy, but as stated, the content of
the waste stream will vary the stoichiometric ratio due to the quantity of oxygen proximately present in the waste stream. The addition of energy in [2] is described by the
symbology . The added energy is from the plasma torch used to motivate the
reaction, and is captured in the conservation of energy employed in the heat balance.
The equilibrium assumption is verified from a well-documented application of a system
of independent equations commonly used in the analysis of gasification reactions (36),

20

As noted by Zhang, et al., the detailed pyrolitic process of gasification for such a
complex feedstock as municipal solid waste is not well understood (38). The vitreous
byproducts were thus assumed. In this study, it is recognized that the presence of these
inorganic constituents will serve to increase the required plasma arc energy requirements.
21
The value for m used in this analysis is 0.47 mol/mol of waste in accordance with the
study presented by Mountouris, et al. (27), as it produces the most hydrogen for the least
energy.
31

and is reasonable because of the high temperatures involved . This follows from the
application of the phase rule of chemical equilibrium that relates the number of coexis tent phases and the degrees of freedom (independently variable environmental conditions
controlling the physical state of the components) to the components that comprise those
phases.
The solution of the global gasification reaction will require seven independent equations to solve for the seven unknowns (n 1, n 2, n 3, n 4, n 5, n 6, and n 7). The first three of
these independent equations will be determined from the water-gas shift, the methane
decomposition reaction, and the steam-carbon reaction. The enthalpy balance gives the
fourth equation, and the partial mass balances of the products and reactants provide the
last three.
From the phase rule and an understanding of the compounds of the products and reactants of equilibrium, a subset of gasification reaction equations are determined (27).
These equations are used to establish the reversible composition of methanea process
noted to be favored in gasification modeling in moisture rich environments 22 (given by
[3]) (37), the catalytic water-gas shift (given by [4]), and the endothermic steam-carbon
reaction given by [5].

[3]

[4]

[5]

22

The moisture rich gasification process is also favored in this study for the understan ding of waste streams of significantly wet feedstock as present in a primarily organic
waste stream.
32

It is important to discuss the subject of chemical equilibrium on which this analysis


is based. The simplifying assumption of chemical equilibrium is based on the observation that for high temperature reactions, equilibrium is quickly attained. At the typical
temperatures of plasma gasification 1200 K 2300 K, the reactants are quickly raised to
a transition state (the point at which chemical and molecular bonds are severed) very
rapidly 23 (38).
The analytical method for determining the equilibrium state of the gasification reaction proceeds by employing the relationship of Gibbs free energy to the equilibrium
constant of the reaction as given by (39) (40)

[6]

where

is the standard free-energy change (in kJ/kmol),

stant (kJ/Kkmol), and

is the ideal gas con-

is the temperature of the reaction in kelvin.

is the equilibrium

constant of the reaction equation which is computed from the law of mass action (which
relates the concentrations of the reactants to the rates of reaction) for each of the governing gasification reactions that comprise the overall gasification reaction24. For instance,
in the case of the endothermic steam carbon reaction [5], the equilibrium constant is
given by

[7]

Then, utilizing the definition of Gibbs free energy applied at constant temperature and
pressure (41)

Residence times are reported in the plasma gasification process on the order of 10s of
seconds, and the assumption of molten inorganics in the process provides good indic ation that equilibrium is a valid assumption under these conditions.
24
Details of this determination are standard in chemical engineering and are referenced
here from the Handbook of Environmental Engineering Calculations (22)
33
23

[8]

the standard free-energy change is usefully related to the change in enthalpy (

taken here as the difference between the heats of formation and the heats of reaction of
the compounds in the reaction), the gasification temperature, and the change in entropy
(

which conveniently provides a relation to the energy absorbed from the enviro n-

ment as the process progresses towards thermodynamic equilibrium).

[9]

Combining the relationships yields a useful formula for the determination of the energy
balance for the overall gasification reaction

[10]

Noting that taking the partial of [8] with respect to temperature in a constant pressure environment yields

)
[11]

and presenting this in a more useful form by performing the differential with respect to
temperature reveals

34

)
]

[12]

and introducing [8] and [11] provides the Gibbs-Helmholtz equation (41)

)
]

[13]

which is central in relating the temperature to equilibrium for a constant pressure env ironmental process. It is this relationship that is used in conjunction with the Gibbs free
energy and at a constant pressure reaction; we have the relationship from a substitution
of [6]

[14]

The equilibrium constant is thus evaluated by the application of the heat capacity
correlation (22)

[15]

35

where the coefficients A, B, C, D, and E are the regression coefficients for the chemical
compounds from the empirical tables. Integrating [15] at constant pressure for the
change in thermal states of the reaction gives

]
[16]

reintroducing this to [14] gives (27)

)
[17]

where

and with the values of A, B, C, D, and E representing the

change in species coefficients for each of the equilibrium reactions. The values T1 and
T2 are the atmospheric temperatures and the gasification temperatures, respectively.
Using again [5] the endothermic carbon-steam reaction, as an example,

[18]

In addition to the above independent equations used to determine the three independent equilibrium constants (that is the application of Equation [17] to Equations [3],
36

[4], and [5]), four additional equations are required to satisfy the number of unknown
coefficients in the overall gasification reaction. To do this, we employ the enthalpy
balance of the overall gasification reaction (39). Relating the heat of formation of the
products to the heat capacity of the reactants and the interjected energy is the relatio nship provided by [8] and, along with [9],

[19]

and introducing [15] (39)

[20]

with the products and reactants of the overall gasification reaction, accounting for the
entropic term as energy input from the plasma arc (Q) to motivate the reaction gives the
heat balance for the gasification reaction (27)or our fourth independent equationas

(
)
[21]

It should be noted here that the heats of formation (Section 7.4.3) of the gaseous compounds and water (liquid and vaporous) are the terms ,
(Section 7.4.1) are given by,

, and the heat capacities

, are taken from Yaws (22). The chemical properties


37

of this reference are used due to the analogous nature of the thermal characteristic within
the temperature bounds of the gasification reaction used in this analysis (285 K 1500
K), with the exception of solid carbonthe soot termhaving a maximum applicable
temperature of 1100 K. This ensures the analysis operates within the bounds of the
measured parameters.
To solve the system of simultaneous equations three additional relationships are
needed. These proceed from the partial mass balances of carbon, hydrogen, and oxygen
(the three elements present in the gasification reaction modeled) from the gasification
reaction [2] (27):
the carbon balance,

[22]

the hydrogen balance,

[23]

and the oxygen balance,

[24]

Using a numerical solver for the seven equations (equations [3], [4], and [5] whose
equilibrium constants form the independent equations as determined by equations [17]
and [18], the enthalpy balance given by equation [21], and the partial mass balances
given by equations [22], [23], and [24]) and seven unknowns (n 1, n 2, n 3, n 4, n 5, n 6, and
n 7, the coefficients of equation [2]) yields the values of the coefficients of the chemical
reaction, from which we determine the molar concentrations of the products, reactants,
and the input energy. The detailed methodology of the calculations is documented in a
routine written in Mathematica that allows for multivariate analysis , determining the
38

significant contributors to the gasification reaction and calculating the energy consumption of the plasma torch. The primary Mathematica routine is documented in the
Appendix Section 7.2.

3.3 Syngas conditioning


From the gasification process, the byproduct of most interest for th is evaluation is
polluted. The syngas created contains not only hydrogen (useful for fuel cell induction
for the production of electrical power) but also nitrogen, carbon monoxide, carbon
dioxide, and trace gases such as methane, hydrogen sulfide (sour gas), and gases containing volatilized metals. To make the synthesis gas more useful, it needs to be
conditioned. A minimum amount of conditioning may be employed for fueling combu stion oriented processes, as the most environmentally damaging species of gases are
existent in only small percentages in the syngas resulting from plasma gasification (42).
Also, the large percentage of CO can also be combusted. However, of interest in this
study, is the purity of hydrogen for induction into a fuel cell. Commercial grade fuel
cells require high purity hydrogen to levels of quality of 99.99%25. To attain levels this
pure, significant syngas treatment is required.
The model for syngas treatment for this study follows the process evaluated by
Byun, et al. (42). The synthesis gas is cooled, passed into a bag filtration system for the
removal of particulates and volatilized metals, followed by a wet scrubber for the
elimination of acids. The cleaned syngas consisted of the primary species of this study
(H2, CO, CO2, H2O, and N2), with trace amounts of CH4, and H2S. To attain separation
of hydrogen, the syngas is introduced into first a water-gas shift (WGS) reaction 26 (see
25

Commercial grade hydrogen is commonly regarded per SAE J2719. This gas quality
standard has widespread adoption for use in fuel cell vehicles and other commercial
solid oxide fuel cell (SOFC) usages. It is noted that proton exchange membrane (PEM)
fuel cells require higher purity H2 to prevent membrane COx contamination.
26
As noted in Section 3.2.3, the water-gas shift reaction is not the application of a new
technology, but an industry practice that is commonly used in gasification reactions.
Much research has been noted that aims to improve catalyst effectiveness and lower
reaction temperatures, aimed at product aimed at reducing cost and complexity. However, the cycle mimicked in this analysis is industrially proven and of demonstrated
capacity.
39

[4]) for the increased production of hydrogen from carbon monoxide and water. Byun,
et al (42) performed this in two stages for more efficient production of H 2 by allowing
for a high-temperature shift reaction using an iron-chrome catalyst and a lowtemperature shift reactor using a copper-zinc catalyst 27. Between the two stages, the
syngas is cooled and hydrogen sulfide (sour gas) was removed using a fluidized bed
reactor, serving to further purify the syngas, but also to ensure t hat the copper-zinc
catalyst was not polluted by the aggregation of sulfur.
Assuming the water-gas shift reaction occurs at chemical equilibrium, the coefficients of the species in the reaction are scalable. Understanding this, and using the
empirically determined relationship noted by Byun, et al. (42)28, a species production
ratio is determined as the outlet syngas composition relative to the entrance syngas
composition. This production ratio is applied directly to the syngas that is generated
from the process described in Section 3.2.3. The related H2S removal is assumed to
occur similarly.
Table 3-2: Water-gas shift reaction as experimentally demonstrated by Byun, et al (42). The calculated
production ratio is used for hydrogen reformation in the distributed gasification analysis presented h ere.

Syngas Composition
CO
CO2
H2
N2
H2 O

27

Inlet
Average

Outlet
Average

Percentage
Change

Production
Ratio

38.8%
14.5%
36.5%
7.3%
4.0%

0.3%
37.2%
51.2%
7.6%
4.0%

-99.2%
157.4%
40.3%
4.4%
0.0%

0.00773
2.57400
1.40329
1.04396
1.00000

The water-gas shift is exothermic, with the increase in temperature being provided by
the reaction itself as it is motivated by the catalyst. Only sufficient pressure is required
in the operation of this reaction.
28
Noting that the average compositions do not exactly equate to the full composition of
syngas, there are probable errors introduced at this stage, but the errors are estimated to
be within 1-3% of the species by volume, and are acceptable for the gross evaluation
presented in this analysis.
40

Following the water-gas shift, the syngas, rich in hydrogen and carbon dioxide, is
introduced into a pressure swing adsorption (PSA) unit 29. This unit functions by introducing the gas into tank containing a molecular adsorbent that, when under sufficient
pressure, allows carbon-dioxide to embed in the adsorbent. The gas stays contained in
the tank for sufficient residence time, and is then evacuated. In the evacuated process,
the CO2 is released and can be off-gassed or stored for use in the bagged filtration step
where it is used to free particulate matter from the sides of the filter. This process can be
repeated to the maximum efficiency of the adsorbent, and exiting the pressure-swing
adsorption process is a purified hydrogen gas. Hydrogen gas purities have been empirically measured following such a process exceeding that required for commercial grade
usage per SAE J2719 (42). The hydrogen purification process described is depicted in
Figure 3.5.
For the purposes of this analysis, efficiencies of the syngas conditioning cycle are
assumed, as it is beyond the scope of this study to select catalysts and adsorbents for
system optimization for the production of hydrogen from WGS and PSA, though this in
future work could significantly improve the results presented.

3.4 Energy production and storage


The final components in the process are those required for the creation of electrical
power from the purified hydrogen syngas and for electrical power storage. The component that allows for the entire process is one of the more significant consumers of
electrical power, the plasma arc system.

3.4.1 Plasma arc system


The plasma arc system is arguably the central component in the waste gasification
cycle. As noted in Section 2.1, the application of plasma gasification is not new technology, but the application of plasma gasification to solid wastes for the production of
power is a novel undertaking. A variety of plasma gasification systems are in development. Research indicates that both torch type and arc type p lasma systems are being
29

Pressure swing adsorption is a process that requires pressure and evacuation, valving,
and computer or mechanical controls, but no significant external energy input.
41

built, tested, and evaluated. For this analysis, the assumed system is a transferred arc
gasification system, depicted in Figure 3-6.
The advantages of the transferred arc units are increased efficiency (between 65%
and 85% of the power consumed by the arc is transferred to the working material, as
opposed to 60%-75% of that of a plasma torch) (43), and improved electrode life. The
detriments include difficulty of maintenance (replacement of the electrodes is more labor
intensive than the replacement of a simple torch assembly) and the increased cooling
requirements as there is no accelerated working fluid (such as steam or air) t o assist in
the preservation of the electrodes, as there is in a torch based system. Some research
suggests additional benefits in torch-based gasification systems that are not accounted
for in a transferred arc configuration. The pressurized gas that is fed through the torch
tip exits at sufficient stagnation pressure as to motivate the waste product as it is being
molten in the plasma furnace (42). In a transferred arc system, the agitation of the waste
product (if necessary) must otherwise be accomplished.
The author believes that there are advantages of a novel arrangement of transferred
arc plasma that will capitalize upon the benefits of efficiency, while allowing for a

Figure 3-5: Hydrogen reformation process from syngas, showing the system inputs to providing for the
water-gas shift reactions and the pressure-swing adsorption system.

42

compact reactor and crucible in the plasma chamber. The basis is to run a series of
plasma arcs through the waste particulate as it rests inside the crucible, similar to methods used in induction furnaces for the large scale smelting of metals. Such a method
may not require cooling 30, but it does require the invention of a cathodic crucible, as the
waste stream is primarily non-conductive. In this arrangement the plasma arcs are
generated from a graphite anode and directly attack the crucible with the introduced
waste being volatilized in the process. Research has revealed similar apparatus used in
laboratory environments (44), though this concept will require some invention .
Because this analysis uses plasma arcs instead of gas or steam torches, transfer efficiencies of the plasma reactor above 80% are assumed (43). To improve durability of
the anode and the crucible in such an arrangement, the working chamber may be evac uated prior to the introduction of power, however system complexity would escalate as an
accelerant would need to be introduced in the chamber to allow stoichiometric equilibrium for the gasification reaction (recall in Section 3.2.3 the system is based as an open
environment with excess air). This study does not introduce these complexities. Instead, the system described in Figure 3-6 is simple. Waste is shredded once a target
input weight of waste is present in the hopper. Conveyors lightly compress the RDF and
introduce it to plasma chamber. The chamber closes and seals, allowing then only for
the air present in the chamber to contribute to the gasification process. In this way,
excess air may be controlled by the waste displacing a known quantity of airin other
words, the plasma chamber would be sized based upon the target input weight to provide
the nominal stoichiometric fuel-air ratio. As the plasma arcs are only energized after
waste is input, the system conserves electrical power, using only that which is required
to gasify the waste.

30

Plasma torches are typically run with steam or an oxidant that serves to lengthen the
life of the electrode (43).
43

It is important here to also revisit the prior assumption that the crucible is free of
slag, as it is purged at the end of each cycle, with the syngas exit tube being valved shut
(the syngas is evacuated at the end of the gasification process), and pumped air being
introduced. The slag is not separated, but allowed to cool into bricks outside the gasification chamber. Because the vitrified solids and metals separate by buoyancy when
molten and because of the solidification rate when heat is removed (45), the cooled brick
can be easily recycled. This system requires timing and sensing logic and is envisioned

Figure 3-6: Plasma arc gasification system as concepted for small scale gasification. The process
runs only until the RDF is completely consumed, allowing for small amounts of waste to be
processed discontiguously and efficiently.

44

to be completely robotic.

3.4.2 Fuel cell


Due to the lack of damaging emissions, robust operation with minimal maintenance,
and efficient packaging, the fuel cell is the electrical powerhouse of choice for this
study. There are, however, a multitude of fuel cell types that are conducive to use in the
process defined. Operating with electrical efficiencies of greater than 40% and due to
the long operation life with minimal maintenance, the proton exchange membrane or,
more accurately, polymer-electrolyte membrane (PEM) type fuel cells are chosen for the
process (46). This type of fuel cell is conducive for uses in automotive applications:
the low temperature operation of the PEM fuel cell reduces the cost of maintenance due
to improved material durability, complex cooling systems are not required, and the cells
result in nearly instantaneous31 production of electricity upon the attained fuel gas and
pressure conditions being satisfied (47). There are noted disadvantages to PEM-type
fuel cells, most significantly the cost of production due to the substantial presence of
noble-metal catalysts (46). The type of catalyst also promotes carbon monoxide poisoning requiring high purity introduction of hydrogen, which for this example is satisfied by
the hydrogen reforming process described in Section 3.3.
A fuel cell commercially available for automotive applications is selected as part of
this study. The PEM cell requires hydrogen to be delivered at 16 bar, at a rate of 2.5
grams/second. As noted in Section 3.3, the hydrogen required is commercial grade with
a purity of 99.99%. In exchange, the cell produces 150 kW of stable electrical power at
320 amps)32. The selection of this cell is based upon the assumed community size and
waste stream to at least recover sufficient electrical energy to power the gasification unit.
The solid-oxide fuel cell (SOFC) is worthy of mention in the selection of an electrical power plant for stationary, distributed power applications. They are among the
highest efficiency fuel cell configurations available, boasting electrical efficiencies
31

Dynamic modeling and experimental verification of a PEM fuel cell shows stabilization of current and voltage within 4 seconds of introduction of sufficiently supplied fuel,
and stabilization of response within 0.1 s econds for a stepped current draw (47).
32
The fuel cell selected is the Ballard FCVelocityHD6 (51).
45

greater than 60%, prior to any form of waste energy (heat) recovery. The benefits of
heat recovery from a solid-oxide fuel cell are also very significant. As the fuel cell stack
may operate at temperatures nearing 1280 kelvin, the thermal efficiency of heat transfer
promotes heat recovery (co-generation) that boosts overall system efficiencies above
80% (46). The high-temperature operation also encourages the development of on-board
reformation of the input fuel, eliminating the need for external facilities for highpurification of the input fuel. They are also resistant to concentrations of sour-gas and
are not influenced by the presence of carbon monoxide. Unfortunately, it is the same
high temperature operation enabling these properties that is the main detractor for solidoxide fuel cells for the distributed gasification process noted here. The high temperature
requires significant complexity in stack cooling and expensive alloys for durability
within the operating environment. Materials also offset the reduced cost of the fuel cell
stack, making SOFC technology expensive for implementation. As SOFC technology
advancements continue to reduce costs (or for significantly larger applications where
costs may be more offset by economies of scale), solid-oxide fuel cells should be considered for distributed power generation resulting from the production of syngas.

3.4.3 Waste feeding system


Battery storage is among the final considerations of this analysis. As the process
described is targeted for local community usage on a distributed basis, the storage of
electrical power generated from the waste gasification unit is important to recover power
for the gasification process allowing it to be at least energy neutral. To understand the
importance of this point, a macroscopic review of the proposed gasification facility
needs to be described.
Waste in this model is gathered at a location close to the point of generation primarily by human means to reduce the energy requirements of waste collection an d
transportation. Such a method of waste collection is feasible in condensed community
environments such as tenements, campuses, businesses, hospitals, and small residential
communities. On such a small basis, however, large volumes of waste will not be
introduced to the gasification cycle on a continuous stream, as is possible with large
46

scale municipal waste treatment. Instead, an advantageous method of introducing the


waste is to start the cycle once a designated quantity of waste is available for pro cessing.
The concept can be executed by having a mass based induction system, as shown in
Figure 3-6 and shown in the system block diagram in Figure 3-7, that begins with
powering a waste shredder. The waste shredder reduces the waste into particulate that
allows it to be introduced into the plasma gasification unit through a waste transfer unit.
Electrical power is then used to initiate the plasma arc for the creation of syngas 33,
previously detailed in Section 3.2.3. The syngas is delivered at pressure (requiring pump
power) into the syngas processing unit outlined in Section 3.3, and it is also pumped into
the fuel cell (to meet the pressure requirements of the PEM-type fuel cell system described in Section 3.4.2). Power generated from the hydrogen is battery stored for future
usage. Commercially available systems for processes requiring electrical power are
typically operated on alternating current. To satisfy this, a D/C inverter (which itself

Figure 3-7: Block diagram detailing the concept of mass-based waste induction for the distributed
gasification of municipal solid waste.
33

Due to the presence of residual matter and to ensure complete gasification, the cruc ible of the plasma reactor should be preheated. This pre-heating creates a molten slag
prior to the introduction of additional waste, serving to aid in vitrification of inorganics
and complete gasification of organics (60). The electrical power requirement for the
preheating of the slag is not accounted for in this analysis , as it is assumed there is no
residual matter from the prior cycle.
47

consumes some electrical power) is needed. Any excess electrical power could then be
used for community purposes. For this process to work as planned, battery storage is
important.

3.4.4 Electrical energy storage


There are various properties that are considered in the selection of an electrical sto rage media. Primarily, the size of the system in electrical capacity, allows for the down selection to a subclass of electrical storage systems. The information presented by the
United States Department of Energy provides a high level review of commercially
available grid-based electrical storage options ordered on system capacity, summarized
in Figure 3-8.

Figure 3-8: Grid-based electrical power storage technologies depicted by discharge potential relative
to system capacity. This information is presented from the United States Department of Energy
Sandia National Laboratories (48).

Due to the cyclic operating nature of the mass-based induction gasification system,
continuous cyclic performance is required. Power generated and used in the system
48

occurs at relatively rapid rates of charge and discharge. In plasma gasification processes, wastes have been converted to syngas at rates of 50kg/hour (42), the time for the
water-gas shift and pressure-swing adsorption reactions is on the order of a few minutes
for flow rates on the order of a 100 liters per minute of syngas (42), and the operation of
the fuel cell stack is also nearly instantaneous, as described in Section 3.4.2. These
process times require the ability to quickly charge and discharge the electrical storage
system, setting the stage for solid-state battery systems. Lithium-ion batteries are
selected not only due to their ability for rapid charge and discharge 34, but also for their
small packaging characteristics, thermal operating range, partial charge acceptance,
systems simplicity, cyclic endurance, and reasonable cost of implementation 35. Based
upon research conducted by Qian, et al. a controller and inverter couples with a state-ofcharge (SOC) and state-of-health (SOH) battery management system (BMS) for gridbased power storage provides an overall storage efficiency of 92.6% (49). Charging
times were also recorded to be within the range of of an hour to charge a 1kW lithiumion from 30% to 70% SOC.

3.5 Ancillary systems


There are various systems that are not detailed in the prior discussion that are wo rthy of mention, as these systems play crucial roles in the execution of the gasification
process and also require of the system process energy. These systems include the
necessary automated computerized controls for mass flow and process health monitoring
within the plasma gasification and hydrogen synthesis processes, the mechanical waste
processing of the shredder to reduce the waste stream particulate size, and the D/C
inverter to convert the electrical power generated by the system into its more comme rcially useful alternating-current form. Commercially available mechanical and electrical
34

Though cyclic charging and discharging are tolerable, computer controls should be
implemented to ensure deep-discharging does not occur as dendrite formation is accelerated, promoting storage decay and shorting (49).
35
Uses of lithium-ion batteries in automotive applications challenges the production of
an inexpensive vehicle, but considered on the kilowatt scale for other electrical power
storage options, lithium-ion batteries are cost competitive when rated against similar
capacity electrical storage options (48).
49

systems were chosen for the anticipated electrical capacity. The energy requirements of
computer controls were not accounted for, but are understood to be of fractional contributions to the electrical requirements of the system.
A mechanical solid waste shredder, capable of grinding and shredding organic and
hard inorganic matter (metals, plastics, and glass) was selected. The durability and
capability of this machine is of key importance. Though it is recommended that glass
and metals are completely diverted from the waste-stream feed, it is reasonable to
anticipate waste stream contamination36. The durability and robustness of the system
should be unencumbered by the accidental (or purposeful) introduction of these components in the waste stream. A unit was selected of the appropriate size class (360
kg/hour). Its operation is driven by a 5 Hp electric motor operating at 20 amps at 230
volts (50).
The envisioned system of waste delivery utilizes a horizontal, augur driven, tubed
conveyor. The benefit of such a system is that the particulate waste can be further
compacted and delivered into the plasma furnace in a single process (with the size of the
furnace chamber and crucible of sufficient mass to accommodate the waste input. Most
of the commercially developed tube-conveying systems are sized for thousands of
pounds per hour throughput. These are too large for a community sized application,
however. The power requirements to deliver and compress the waste are therefore
calculated. The details of the calculation are given in the Appendix Section 7.7.
Concluding the significant power requirements of the gasification process, p umps
are needed for the pressurization requirements of the water-gas shift system, pressureswing adsorption, and for pressurization of the fuel cell. The selected automotive-grade
fuel cell (51) requires pressure delivered at 16 bar(g), or 1,600 kPa. At 230 VAC, a 15
Hp high-purity gas pump is selected to deliver flow at rates of up to 7.25 m3/sec for the
fuel cell at a current of 37 amps. This same pumping power is assumed to be duplicated
for operation of the gas cleaning systems .

36

As expounded upon in Section 7.5.


50

4. Results and Discussion


4.1 Plasma gasification model validation
Because the plasma gasification system is central to the process, the results of the
gasification modeling system are first discussed. The Mathematica program was
generated as formulated in Section 3.2.3, and vetted against the GasifEQ (27) routine
presented by Mountouris. The results show insignificant differences, within solver error,
between the two routines when provided the same input waste stream and ambient
conditions.
Some conditions were modified based upon discussion in prior sections of this do cument. The atmospheric conditions are assumed to be based on the averag e daily
temperature of the New England area of the United States , as stated in Section 3.2.3.
This sets the ambient temperature to 283.7 K [51F], important for the thermochemical
energy balance. All gas conditions outside of the gasification process and syngas
conditioning cycles are evaluated at standard day atmospheric conditions for mass to
volume relationship. The concentration of air is also held steady at 0.47 kmol/kmol
RDF, also noted in Section 3.2.3, to yield a stoichiometric ratio that benefits hydrogen
production while reducing energy consumption (27) (35). This is verified by varying the
molar air ratio while holding all other values of the gasification reaction fixed. As noted

Figure 4-1: Trends of electrical consumption (left) and hydrogen production (right) for varying
molar air ratios

51

in Figure 4-1, increasing molar air ratio decreases power consumption but also decreases
hydrogen production. To find the optimum, the trend of electrical consumption was

Figure 4-2: The trend of electrical consumption (normalized) and hydrogen production (inverted) to show optimum molar air ratio for a given waste stream.

normalized and hydrogen production was inverted. This shows an intercept in the two
trends at 0.44, shown in Figure 4-2. Note, however that the value used in the analysis
presented here is not changed from what was previously assumed. The reason for this is
two-fold, the species ratios of compounds in the waste stream, which vary significantly
in this study, influence the optimum air ratio substantially, and the system desc ribed in
Section 3.4.1 would have a fixed air to fuel ratio by design. It is noted that improved
efficiency can be attained by modifying the air ratio based upon the introduced waste
stream, and further study would be required to ensure the added complexity trades
favorably to generated power.
Appended to the Mathematica code is the Boie estimate for the heating value of
the waste stream. As discussed in Section 3.2.2, the Boie method compares to research
within 4% of the reported heating values of municipal solid waste and similar organic
feeds (34), though it generally overestimates HHV (33). To understand the significance
of 4% on the HHV of the RDF, the Boie estimated HHV was scaled by 0.96. The result
52

was a 14% increase in the power requirements for the plasma torch, with no ch ange (as
expected) on the production of hydrogen. On such a small system as is evaluated here,
such error is significant, however it is noted that the system can still yield supplemental
electrical power for the community with the addition of more RDF. In the practical
application of plasma gasification technology the gasification temperature could be
controlled real-time to minimize power consumption.
The benefits of using Mathematica to model the plasma gasification system are
revealed in the ability to have many multivariate cases processed, aiding in the development of key trends that promote understanding the desired waste stream, arc energy
requirements, and community sizing for the commissioning of distributed plasma arc
gasification systems. For the results presented here, the Mathematica model was
exercised to vary gasification temperature within the bounds of the chemical properties
values that comprise the gasification model (1200 K 1500 K [1700F 2240F]). The
RDF feed was varied by using the typical waste stream proximate analysis provided in
Figure 3-4, but modified to simulate the effect of diverting a percentage of the waste
stream from the gasification process. This was done to mimic a user recycling po rtions of the waste stream prior to introducing it to the gasification process. For each
major category of the waste stream content (e.g. organic matter) the percentage of the
constituent that becomes RDF was randomly varied using a uniform distribution 37.
Limits were placed on the randomized recycling percentages to firstly insure that some
waste content was present in every tested waste stream, and secondly to better mimic
recycling contamination. Finally, the program was allowed to run 100 of the multivariate points per temperature condition to capture significant trends.
Among the first characteristics observed was the significance of gasification te mperature on hydrogen content of the syngas, shown in Figure 4-3. This slightly declining
trend indicates that lower gasification temperatures are desirable, but the temperatures
required to ensure complete vitrification of solids and reductions of exitin g solid particulate (soot) warrants the gasification temperatures be held above 1273 K (27). Doing so
37

i.e., the proximate composition of the typical municipal solid waste stream was fixed
per Figure 3-4, but the recycling percentages were allowed to vary randomly per a
uniform distribution.
53

Figure 4-3: Hydrogen content of the syngas from the plasma gasification furnace as related to the
gasification temperature for a large sample of randomly diverted waste stream constituents

results in 1% less hydrogen content in the syngas (on a volume basis), and a corresponding 14% increase in plasma arc power requirements, as calculated from Figure 4-4.
Understanding the change in hydrogen as a function of the waste stream proximate
content is also useful. When reviewing the data from this perspective, the results may
appear counterintuitive. If so, it is likely due to familiarity with pyrolitic processes in
which optimization is heavily reliant upon calorific performance of the constituents.
This is not the case for the desired creation of hydrogen gas. Figure 4-5 shows the
volume percentage of hydrogen resulting from the concentrations of various constituents
in the waste stream. A steep positive slope of the curve indicates a strong and beneficial
relationship of that constituent to the generation of hydrogen. Similarly a steep negative
slope indicates that the presence of a constituent may detract from the production of
hydrogen.
Textiles provide the greatest hydrogen output. Two things are important in this result: the assumption was made that textiles would include rubber as a means of
54

Figure 4-4: Plasma arc power requirements as a function of gasification temperature for a large
sample of randomly diverted waste stream constituents

simplifying the number of constituents in the waste stream (reference Section 3.2.1), and
the typical content of textiles and rubber is a small percentage of the overall waste
stream. While it is beneficial to have these compounds present, they are not present in
the volumes to produce sufficient syngas alone.
Plastics (as may be expected by reviewing their hydrogen rich chemical co mpounds) have nearly as much influence as textiles. What may not be expected is that
what follows are the organics (food wastes and yard clippings). Food wastes and yard
clippings are high in hydrocarbons, and the high moisture content favors the water-gas
shift portion of the reaction, boosting hydrogen output.
Conversely, wood and paper have negative trends in the production of hydrogen.
Paper is primarily carbon, as is wood, and are relatively dry. Their abundance of carbon
was hypothesized to drive the steam-carbon reaction, equation [5], but their dry nature
does not provide sufficient water to favorably advance the reaction.
55

Figure 4-5: Trends of the concentration of various waste stream constituents on hydrogen gas
production for a large sample of randomly diverted waste stream constituents

In Figure 4-5, it seems that the presence of glass and metals has been neglected
when in fact they have been removed to simplify the graphic. These constituents provided a slightly greater than 0 slope, which indicates the production of hydro gen is not
significantly affected by the presence of glass and metal in the waste stream. The power
requirements of having glass and metals present (even at small percentages) outweigh
their hydrogen performance and should be diverted from the plasma gasification process
(though there is no other detrimental effect with having them present other than the
reduction in net energy generated).
The species present in the synthesis gas as a result of the plasma arc process are also
important. As may be expected from having run the random multivariate analysis over a
range of gasification temperatures and recycling percentages, post-recycling RDF
mixture averaged over all the cases run appears very similar in proximate content to the
typical waste stream. The composition of the syngas is therefore similar to what may be
expected from introducing an unmodified waste stream into the plasma gasification unit.
56

Table 4-1 shows the average syngas composition as given from the global gasification
reaction, equation [2]. There is almost no methane, indicating the weak influence of the
methane decomposition reaction, equation [3], but hydrogen is abundant in the syngas.

Table 4-1: Average composition of synthesis gas from the multivariate plasma gasification analysis.

Volume of Gas Produced by Gasification


H2
(%vol)

CO
(%vol)

CO2
(%vol)

H2O
(%vol)

15%

17%

8%

15%

CH4
(%vol)

0.0001%

N2
(%vol)

45%

Understanding the trends discussed, one final relationship is needed to begin the
macroscopic systems analysis: the post-recycling content of the waste stream. Three
cases were evaluated based upon the results. The program was run: with a waste stream
containing only plastics, organics, and textiles (as a perfect case); with a waste stream
that does not divert any waste from the plasma reactor; and a waste stream that impe rfectly diverts glass and metals 38. The results of this case, graphed in Figure 4-6, show
the most beneficial waste stream is the one that recycles only metals and glass. It has the
lowest power requirement and the lowest required waste generation rate to yield a target
amount of hydrogen production.

38

The imperfect diversion of glass and metals is a contamination rate of 5%. See
discussion in Section 3.2.1.
57

Figure 4-6: The waste generation rate and the plasma arc power requirements to yield 2.6
grams/second of hydrogen in the resulting syngas when three unique waste streams are subjected to
plasma gasification.

4.2 Sustainable system sizing


The sized system begins with the trends developed from the gasification model.
With the gasification temperature set at 1275 K, molar oxygen ratio held constant at
0.47, and the waste stream tailored by diverting glass and metals (excepting contamination), only the change in RDF feed rate affects the production of syngas and the power
required. Developing these trends establishes the fundamentals used to size the system.
The first trend that is observed is that RDF feed rate, hydrogen gas generation rate,
and power consumption are all related linearly, as shown in Figure 4-7. This trend is
important. It shows that the plasma arc gasification process is directly scalable. It also
provides for simple determination of the size of the system. The linear relations among
the RDF feed rate, the gasification power requirement, and the amount of hydrogen
produced (in Figure 4-7) allows a useful determination: by setting a waste feed rate, and
58

reading the power requirement from the upper diagram of Figure 4-7, and reading from
that power requirement to the hydrogen generation rate on the lower diagram of Figure
4-7, one finds that one kg of waste per day produces approximately .0007 grams of
hydrogen per second [1 lb/day RDF .0002 lb/s H2] exiting the syngas reforming
system following the plasma gasification cycle.

Figure 4-7: Feed rate versus power consumption (upper diagram) and hydrogen generation rate
following gasification cycle and water gas shift (lower diagram) showing the linear relationship
among the three paramaters. The waste stream has a contamination content of glass and metals due
to imperfect separation. The hydrogen flow rate is that leaving the syngas reforming system.

59

Sizing the system solely for electrical power production and consumption starts with
the aforementioned relationship. Understanding that the selected fuel cell (51) generates
75 kW of rated power given the introduction of hydrogen at the rate of 1.3 g/s [10.3
lbs/hr], the required power for the gasification cycle to deliver enough syngas to meet
the fuel cell flow rate is 15.2 kW, acting on 1857 kg/day [approximately 171 lbs/hr] of
RDF. This is enough power to fuel the plasma arcs, operating at 80% efficiency, consuming 35 kW 39. Table 4-2 summarizes the system sizing on the basis of electrical
power consumption and hydrogen generation.
The sizing of the system proceeded from the discussion in Sections 3.1 through 3.5,
but it also integrates assumptions not previously stated. Because of the size required for
the system to produce the hydrogen requirements of the selected fuel cell, it is unlikely
that the system will operate intermittently. The power consumption and hydrogen
production therefore is all based on a full 24-hr daily cycle. The facility power requirements (Misc. Facility and Computing as listed in Table 4-2) is assumed to require no
more than the electrical power required for a typical US household (12).
The resulting power produced, 19.6 kW, is sufficient to power 24 residences (12).
The number of residences required to produce waste feed rate needed for that power
ranges from approximately 280 300 typical US households (3).

Table 4-2: Summary of the distributed plasma arc gasification system sizing based upon power
requirements and consumption, showing positive net power production

Efficiency
Shredder (RDF Processing)
Conveyor (Waste Tranfer Unit)
Gasification Unit
Plasma Arcs
Syngas Processing
Fuel Cell
Electrical Conversion & Storage
Misc Facility and Computing
Net Production
39

---80%
--92.6%

RDF Feed
(kg/day)
1857
1857
1857

Pow er
(kW)

H2 Output
(g/s)

-4.6
-1.2
-15.2
-3.0
-25.0
75.0
-5.6
-0.8

--0.93
-0.37
-1.3
--

19.6

The remaining energy goes into producing heat, that is recoverable with at least 30%
efficiency, however heat energy recovery is not regarded in this study.
60

5. Conclusions
5.1 Viability
The study proves viability. Though the size of the unit does not scale to self-power
the individual home or the small community, larger communities (high -rise apartments
and condominiums, hospitals, retail and commercial establishments) and neighborhoods
may see significant benefit from the on-site installation of this type of facility. Supplementary benefits of heat, hot-water, and adsorption cooling could all be pursued to make
the unit more economically attractive. Further in the case of hospitals , especially,
expensive services for the sanitary disposal of wastes may be eliminated (subject to the
regulations for hazardous waste retention and processing). The system can also be
combined with methane based fuel cell systems (or co-mingled fuel cell systems) that
would permit completely distributed power for those locations that already have gas
feeds for commercial or residential services. This would provide the benefits of nearly
uninterruptible power.
Critical to the implementation of the solution is economics. As most business cases
are generated showing benefit for a limited control volume, the case that the proliferation of distributed plasma waste gasification units will eliminate electrical line loss (and
the significant costs associated with it) and prevent the use of resources in centralized
waste collection is a difficult one to argue. Though the costs of commissioning and
maintaining distributed plasma arc gasification systems is not determined in this study,
the cost of articles of the system (e.g. the fuel cell, the battery, the gas reforming cat alysts, and the plasma chamber) undoubtedly makes such a system expensive. To
promote the development and installation of these systems, applications that directly
benefit from zero waste generation, uninterruptible power, remote (grid disconnected
operations) should be the focused areas for this technology.

5.2 Notable errors and future work


Considering the overestimate of the Boie equation for the prediction of the HHV of
the RDF (as noted in Section 3.2.2), the gasification program was run with a 4% error
61

bound (by scaling the value provided by Equation [1]. The difference, as expected, is
present in the computation of the input energy required for gasification. Using identical
inputs, except for the 4% scaled estimate of HHV, provides a 14% difference in input
energy (since the Boie typically overestimates HHV (32) (31), this is a 14% increase in
arc power required). Though this is not enough variation to challenge the fundamental
viability of the system, it is enough variation to warrant the review of HHV prediction
methods with future work possibly employing modifications to Dulong or Boie to
provide more accurate input energy predictions.
Materazzi, et al. (35) have developed in the course of this research a more inclusive
methodology that accounts for the various chemical species that are vitrified in the
plasma reaction of municipal solid waste using a multi-step gasification cycle. This
more comprehensive method may estimate more accurately the energy consumed in the
gasification reaction. Future work on this system would benefit from including M aterazzis thermochemical model, as residual slag heating would be inherently accounted
for, improving the estimate of input energy and quantity of waste to net positive electrical benefit from the system defined in this work. The results of the Matterazzis
research were published only after significant progress on this thesis had been completed.
The research conducted in this study has also revealed significant and growing interest in the topics of distributed power generation, land based fuel cells, battery grid
storage systems, and waste gasification. Efficiencies are ever increasing, and innovative
means of reducing cost and improving durability are continually being discovered. The
coming times may herald the hydrogen and renewable energy economy, based on
distributed power generation. Technologies such as plasma arc waste gasification,
which enable distributed waste elimination and distributed power generation for sustainable community development, have a future in the coming ecological and economic
environment.

62

6. References
1. Scheuerman, William. Globalization. Stanford Encyclopedia of Philosophy (Spring
2011 Edition). [Online] 2011.
http://plato.stanford.edu/archives/spr2011/entries/globalization/.
2. Wilson, Dominic and Dragusanu, Raluca. The Expanding Middle: The Exploding
World Middle Class and Falling Global Inequality. Goldman Sachs Economics
Research Group. New York : Goldman Sachs, 2008.
3. United States Census Bureau. International Programs Total Midyear Population for
the World: 1950 - 2050. U.S. Census Bureau - International Programs . [Online]
June 27, 2011. [Cited: September 6, 2011.]
http://www.census.gov/population/international/data/idb/worldpoptotal.php. IDB
Version: Data:11.0620 Code:11.0615.
4. U.S. Energy Information Administration. International Energy Outlook 2010.
Office of Integrated Analysis and Forecasting, United States Department of
Energy. Washington, DC : U.S. Energy Information Administration (EIA), 2010.
DOE/EIA-0484(2010).
5. International Energy Agnecy. World Energy Outlook 2009 Fact Sheet. Paris :
International Energy Agency, 2009.
6. The World Bank. World dataBank--World Development Indicators (WDI) & Global
Development Finance (GDF). The World Bank--World Databank. [Online] 2011.
[Cited: 02 19, 2011.] http://data.worldbank.org/topic/energy-and-mining.
7. United States Environmental Protection Agency. Municipal Solid Waste in The
United States: 2001 Facts and Figures. Office of Solid Waste and Emergency
Response (5305W). Washington, D.C. : United States Environmental Protection
Agency, 2003.
8. Bogner, J., et al., et al. Waste Management, In Climate Change 2007: Mitigation.
Contribution of Working Group III to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge : Cambridge University
Press, 2007.
9. Environment Directorate. OECD Environmental Data: Compendium 2006-2008 -Waste. Environmental Performance and Infomration Division OECD. Paris :
Organization for Economic Cooperation and Development, 2008.
10. Tchobanoglous, George, Ph.D. Solid Waste Management. [book auth.] Joseph A.
Salvato, Nelson L. Nemerow and Franklin J. Argardy. Environmental
Engineering. Hoboken, New Jersey : John Wiley & Sons, Inc., 2003.
63

11. United States Environmental Protection Agency. Municipal Solid Waste


Generation, Recycling, and Disposal in the United States: Facts and Figures for
2008. Washington, D.C. : United States Environmental Protection Agency, 2009.
EPA-530-F-009-021.
12. U.S. Energy Information Administration. State Electricity Profiles 2009.
Washington, D.C. : U.S. Department of Energy--Energy Information
Administration, 2011. DOE/EIA-0348(01)/02.
13. Fitzpatrick, Richard. The Physics of Plasma. Austin : University of Texas at
Austin, 2008. pp. 6-10.
14. Gasification of municipal solid waste in the Plasma Gasification Melting process.
Zhang, Qinglin, et al., et al. 2011, Applied Energy, p. 7.
15. Higman, Christopher and van der Burgt, Maarten. Gasification. Burlington :
Elsevier Science, 2003.
16. Schecter, Arnold. Dioxins and Health. New York : Plenum Press, 1994.
17. Quapp, William J. General Description of the Plasma Enhanced Melter. Richland :
Integrated Environmental Technologies, LLC, 2002. p. 19.
18. Noyes, Robert. Unit Operations in Environmental Engineering. Norwich : William
Andrew Publishing, 1994.
19. Moustakas, D., et al., et al. Demonstration Plasma Gasification/Vitrification System
for Effective Hazardous Waste Treatment. Journal of Hazardous Materials. 123,
2005, Vol. B, 038.
20. Westinghouse Plasma Coal Gasification and Vitrification Technology.
Westinghouse Plasma Corporation. Hersey : Westinghouse Plasma
Corporation, 2002. Power Generation Conference. pp. 17-18.
21. Cannon, James S. Greening Garbage Trucks: Trends in Alternative Fuel Use,
2002-2005. New York : Inform, Inc., 2006.
22. Yaws, Carl L. Yaws' Handbook of Thermodynamic Properties for Hydrocarbons
and Chemicals. New York : McGraw-Hill, 2009. ISBN 978-1-60119-797-9.
23. U.S. Environmental Protection Agency. Waste Reduction Model (WARM) Version
12. [Software Documentation] Washington, D.C. : s.n., 2012.
24. Kutz, Myer. Environmentally Conscious Materials Handling. Hoboken : John
Wiley & Sons, Inc., 2009. ISBN 978-0-470-17070-0.

64

25. Chang, Shoou-Yuh. Municipal Solid Waste Management and Disposal, in


Environmentally Conscious Materials Handling. [ed.] Myer Kutz. Hokoken :
John Wiley & Sons, Inc., 2009 .
26. Tchobanoglous, George, Theisen, Hillary and Vigil, S. A. Integrated Solid Waste
Management: Engineering Principles and Management Issues. Michigan :
McGraw-Hill, 1993. 0070632375, 9780070632370.
27. Solid waste plasma gasification: Equilibrium model development and exergy
analysis. Mountouris, A., Voutsas, E. and Tassios, D. Athens, Greece : Energy
Conversion and Management, 2005, Vol. 47. 0196-8904.
28. Biomass steam gasification -- an extensive parametric modeling study. Schuster, G.,
et al., et al. Vienna : Bioresource Technology, 2001, Vol. 77. 0960-8524/01.
29. Mason, D. M. and Gandhi, K. Formulas for Calculating the Heating Value of Coal
and Coal Char: Development Tests and Uses. Chicago : Institute of Gas
Technology, 1980.
30. Niessen, Walter R. Combustion and Incineration Processes: Applications in
Environmental Engineering, Third Edition. New York : Marcel Dekker, Inc.,
2002. ISBN: 0-8247-0629-3.
31. Capareda, Sergio Canzana. Sustainable Growth and Applications in Renewable
Energy Sources. [ed.] Majid Nayeripour and Mostafa Kheshti. Croatia : InTech
Europe, 2011. ISBN 978-953-307-408-5.
32. Buckley, T. J. and Domalski, E. S. Evaluation of Data on Higher Heating Values
and Elemental Analysis for Refuse-Derived Fuels. Gaithersburg : National
Bureau of Standards, Chemical Thermodynamics Division, 1988.
33. Siegle, Volker, Spliethoff, Hartmut and Hein, Klaus R.G. Characterization and
Preparation of Biomass for Co-Combustion with Coal. Dallas : American
Chemistry Societly, 1998.
34. Plasma Gasification of Sewage Sludge: Process Develop,ent and Energy
Optimization. Mountouris, A., Voutsas, E. and Tassios, D. 2008, Athens :
Energy Conservation and Management, 2006, Vol. 49. 0916-8904.
35. Thermodynamic modelling and evaluation of a two-stage thermal process for waste
gasification. Materazzi, Massimiliano, et al., et al. 108, s.l. : Fuel, 2012, Vol.
2013. 0016-2361.
36. John Wiley & Sons, Inc. Kirk-Othmer Encyclopedia of Chemical Technology.
Wiley Online Library : John Wiley & Sons, Inc., 2007. pp. 377-403. Vol. 2.
ISBN: 9780471238966.
65

37. Wiley. Wiley Critical Content: Petroleum Technology Volume 2. s.l. : John Wiley &
Sons, Inc., 2007. ISBN: 978-0-470-13402-3.
38. A thermodynamic analysis of solid waste gasification in the Plasma Gasification
Melting process. Zhang, Qinglin, et al., et al. s.l. : Applied Energy, December
2013, Vol. 112, pp. 405-413. ISSN 0306-2619.
39. Swaddle, T. W. Inorganic Chemistry - An Industrial and Environmental
Perspective. San Diego : Academic Press, 1997. ISBN 0-12-678550-3.
40. Thermodynamic equilibrium model and second law analysis of a downdraft waste
gasifier. Jarungthammachote, S. and Dutta, A. s.l. : Energy, 2007, Vol. 32.
ISSN: 0360-5442.
41. Valsaraj, Kalliat T. Elements of Environmental Engineering: Thermodynamics and
Kinetics. Second Edition. Boca Raton : CRC Press LLC, 2000. pp. 28-40. ISBN
1-56670-397-2.
42. Hydrogen Recovery from the Thermal Plasma Gasification of Solid Waste. Byun,
Yougnchul, et al., et al. s.l. : Journal of Hazardous Materials, 2011, Vol. 190,
pp. 317-323. ISSN: 0304-3894.
43. High Temperature Technologies Corp. Waste to Energy and Waste to by-products
plama pyrolosis systems. High Temperature Technologies Canada. [Online]
2009. http://www.httcanada.com/arc.html.
44. Gupta, C. K. Extractive Metallurgy of Molybdenum. Boca Raton : CRC Press, Inc.,
1992. pp. 357-374. ISBN: 0-8493-4758-0.
45. Novel and innovative pyrolysis and gasification technologies for energy efficient and
environmentally sound MSW disposal. Malkow, Thomas. s.l. : Waste
Management, 2004, Vol. 24. ISSN 0956-053X.
46. United States Despartment of Energy. FCT Fuel Cells: Types of Fuel Cells.
Energy Efficiency & Renewable Energy: Fuel Cells. [Online] USA.gov, March 8,
2011. [Cited: April 6, 2014.]
http://www1.eere.energy.gov/hydrogenandfuelcells/fuelcells/fc_types.html.
47. A Dynamic Model of a PEM Fuel Cell System. Lee, J. M. and Cho, B.H.
Washington, DC : APEC 2009. Twenty-Fourth Annual IEEE, 2009. Applied
Power Electronics Conference and Exposition. APEC.2009.4802740.
48. Akhil, Abbas A., et al., et al. DOE/EPRI 2013 Electricity Storage Handbook in
Collaboration with NRECA. Albuquerque : Sandia National Laboratories, 2013.
49. Qian, Hao, et al., et al. A High-Efficiency Grid-Tie Battery Energy Storage System.
IEEE Transactions on Power Electronics. 3, 2011, Vol. 26, March 2011.
66

50. Franklin Miller, Inc. Taskmaster TM8500 Twin Shaft Grinders with Cutter
Cartridge Technology. Franklin Miller. [Online] Franklin Miller.
http://www.franklinmiller.com/taskmaster-8500.html.
51. Ballard Power Systems, Inc. BallardFuelCellSpecs_ProductPortfolio.pdf. [Online]
September 2012. http://www.ballard.com/files/PDF/Bus/HD6.pdf.
52. Krugman, Paul R. and Obstfeld, Maurice. International Economics - Theory &
Policy. Boston : Addison Wesley, 2009. HF1359.K78 2009.
53. International Data Base - Region Summary - U.S. Census Bureau. International
Data Base (IDB). [Online] December 28, 2010. [Cited: January 30, 2011.]
http://www.census.gov/ipc/www/idb/region.php.
54. Collins, Susan V. Contamination in Single-Stream Materials (MRF and post-MRF).
Container Recycling Institute. Epsom, New Hampshire : Northeast Resource
Recovery Association, 2012.
55. Screw Conveyor Power Calculation | Screw Conveyor Capacity Calculation | Screw
Conveyor Formulae. Mechanical Engineering--A Complete Online Guide for
Mechanical Engineer. [Online] [Cited: April 19, 2014.]
http://www.mechanicalengineeringblog.com/2620-screw-conveyor-powercalculation-screw-conveyor-capacity-calculation-screw-conveyor-formulae/.
56. Anthony, W. Stanley and Mayfield, William D. Cotton Ginners Handbook.
Washington, DC : United States Department of Agriculture, 1995.
57. Burnley, S. J. The Use of Chemical Composition Data in Wast Management
Planning - A Case Study. Waste Management. May 2, 2006, pp. 327-336.
58. United States Department of Transportation; Research and Innovative
Technology Administration: Bureau of Transportation Statistics. National
Transportation Statistics. Washington, DC : Research and Innovative
Technology Administration (RITA): U.S. Department of Transportation (US
DOT), 2012.
59. Lee, C. C. and Dar Lin, Shun. Handbook of Environmental Engineering
Calculations, Second Edition. New York : McGraw-Hill Education, 2007. ISBN:
9780071475839.
60. From Waste to Electricity through Integrated Plasma Gasification/Fuel Cell
(IPGFC) System. Galeno, G., Minutillo, M. and Perna, A. s.l. : International
Journal of Hydrogen Energy, 2011, International Journal of Hydrogen Energy,
Vol. 36, pp. 1692-1701. ISSN: 0360-3199.

67

7. Appendices
7.1 Unsustainable middle class growth
In determining the need for a waste-to-energy solution, research was conducted
concerning the worlds electrical energy consumption, the worlds waste production, and
the rate of growth presented by each. These were then contrasted with the worlds
capability to produce sufficient energy or sufficiently dispose of waste with the rate of
growth projected.

7.1.1 Middle class population rise


The economic report by Goldman Sachs entitled The Expanding Middle (2) provided a reasonable basis from which to base the research. In this report Dominic Wilson
and Raluca Dragusanu document two transformative and related trends: power in
spending is shifting from the richest nations to those of predominately middle income,
and the purchasing power of people of a middle class income is overtaking the purchasing trends of the very rich due primarily to the explosive growth of the middle class
population segment.
For their study, Wilson and Dragusanu relate the middle class in the economic terms
of purchasing power parity to settle differences in currency valuations and purchasing
power (52). In the terms of the United States Dollar, Wilson and Dragusanu deem the
middle class population of the world to earn between $6,000 and $30,000, roughly
equivalent to the median incomes of the OECD nations (2). It was in reviewing the
wage range of the middle class to per capita income projections that the burgeoning
growth of the purchasing power of the middle class became evident, especially among
Brazil, Russia, India, and China, and other developing nations a rapidly growing
grouping of countries known in economics by the acronym BRIC, as shown in Figure
7-1. Using the Goldman Sachs expanding middle class data and correlating it with the
United States Census Bureau projections on world population growth shows the significance of the middle class, the staggering growth of this power population segment is
apparent in Table 7-1.
68

When comparing the worldwide growth of the middle class population (Figure 7-2)
an interesting trend is visible. Between the years 2004 to 2008, a significant increase is
seen in the growth rate of the worldwide middle class, with no correlating growth rate
increase in the world population projections. The growth rate of the middle class is even
seen to be similar to the growth rate of the world population for the years between 2010
and 2030.

Figure 7-1: Excerpt from Goldman Sachs report on the ex panding middle class showing a pareto of
the per capita income in 2007 and comparing it to the world projection in 2050. The chart on the
right shows a swell of middle class income, particularly in the nations known as the BRICs.

Table 7-1: The growth of the world middle class compared to the growth of the world
population: a compilation of data from the Goldman Sachs* (2) and the U.S. Census Bureau** (53).

Year
1960
1968
1972
1980
1990
2000
2004
2008
2010
2020
2030

World Population**
3,042,445,344
3,562,353,760
3,867,338,018
4,452,942,594
5,289,040,477
6,089,648,784
6,393,741,245
6,700,983,106
6,852,472,823
7,592,888,345
8,248,535,284

World Population at
Middle Class Lev el*
500,000,000
750,000,000
800,000,000
1,000,000,000
1,010,000,000
1,100,000,000
1,350,000,000
1,550,000,000
1,800,000,000
2,750,000,000
3,600,000,000

69

World Population
w ithin Middle Class
Status
16%
21%
21%
22%
19%
18%
21%
23%
26%
36%
44%

Figure 7-2: World population growth showing the growth of the middle class. Data sources:
Goldman Sachs* (2)and the United States Census Bureau** (53).

7.1.2 Middle class electrical power consumption


It may be assumed that with the increasing middle class popu lation comes an increased electrical power demand. To test this claim, aggregated data from The World
Bank World Databank of Energy Consumption (6) and projections of energy usage from
the United States Energy Information Administration International Energy Outlook (4)
were combined and correlated to the worldwide population and worldwide middle class
projections from the U.S. Census Bureau and the Wilson and Dragusanu report on the
Expanding Middle for the correlating years. Inspection of this data shows that the
trend of the increasing rate of energy consumption matches the rate of worldwide middle
class expansion, even to the same period of its upswing, as shown in Figure 7-3. As the
data in worldwide electricity consumption arrives from two different sources and show
parity for overlapping years, the existence of a trend is more evident, and this trend is
interrogated further by linearly correlating the projection of the worldwide electricity

70

production to the prediction of the worldwide middle class population between the years
2007 to 2030, shown in Figure 7-4.

Figure 7-3: World electrical energy consumption on the basis of the world population. This
chart shows correlation to Figure 6.2, where the increase in the world wide middle class correlates to the growth of the world projections of energy consumption.

Figure 7-4: The projected world electricity consumption against the projected middle class
population growth showing close linear correlation between the global middle class and power
consumption.

71

This analysis is a major tenet to the claim of Figure 1-2 in that it relates the key influencing factor in future electrical demand being a growing global class of people of middle
income. From the same analysis presented in Figure 1-2, it may be argued that there is
no actual correlation to the growth of the middle class and global electrical power
consumption. Refuting this argument is the Wilson and Dragusanu report by Goldman
Sachs indicating that the size and thus purchasing power parity of the global middle
class prior to the year 2004 was such that even large swings in middle class consumption
had a small influence on key global indicators, but the recent and explosive expansion of
this population group is shifting the balance of economic power to the global middle
class. One need only review Table 7-1 to see that by 2010, more than 25% of the
worlds population exists at the middle class level, a trend that grows to 46% only
twenty years later.
The question still remains, can the world support such growing demand for electrical power? Because the natural resources exist to sustain such growth 40, the answer is
largely economic (5). The International Energy Agency (IEA) states that an additional
4,800 gigawatts of electrical power will be needed to support the demand in the year
2030 at the cost of 35 billion dollars per year if capital investments were to have started
in 2008. This is no small investment, and the current rate of infrastructure growth and
improvements worldwide will still leave 1.3 billion people without electrical power in
2030 (12).

7.1.3 Middle class waste production


Applying the law of conservation to the macrocosm of the global middle class ind icates that with increased consumption comes increased waste. The data to support such

The worlds energy resources are adequate to meet the projected demand increase
through to 2030 and well beyond. But these Reference Scenario trends have profound
implications for environmental protection, energy security and economic development.
The continuation of current trends would have dire consequences for climate change.
They would also exacerbate ambient air quality concerns, thus causing se rious public
health and environmental effects, particularly in developing countries. (12)
72
40

an argument however is mired in uncertainty. In fact, Jean Bogners environmental


waste management research team concluded
The availability and quality of annual data are major problems for the
waste sector. Solid waste and wastewater data are lacking for many countries, data quality is variable, definitions are not uniform, and interannual
variability is often not well quantified. (8)
The team continued on to describe industry accepted methodologies for determining
waste output and understanding trends. The methodologies they present (the use of
internationally collected survey data and statistics, estimates on the basis of population
dynamics, and the use of demographic and economic indicators) are employed in this
thesis.
Bogners team provides the first insight towards a potential link between the gro wing worldwide middle class and waste generation rates. In Figure 7-5 and in Figure 7-6
a significant trend is notable: throughout the time scale, the rate of waste generation in
Latin America, South East Asia, North and Sub-Saharan Africa and the Middle East are
significantly higher than the countries in the developed world. Possibly of greater
interest is the step change in waste production in 1994, with a steadily growing trend

Figure 7-5: IPCC unrecycled waste generation data showing minimum annual rates of carbon
storage in landfills from 1971 to 2002. (8) Units of the ordinate axis are teragrams (Tg) of carbon

73

Figure 7-6: PCC graphical representation of post-consumer waste generation between


and 2002 (8). The units of the scales are in teragrams [Tg]

1971

thereafter. This trend appears to roughly relate with the same period as the rise of the
middle class and the previously noted changes in electrical generation. Further illustra ting this trend, Bogner relates the waste generation rates graphically, reproduced in
Figure 7-6. Visually coinciding with the growing middle class is the waste characteristic
in the BRIC nations.
Providing an objective measure to this waste trend is of the most interest in the argument that waste projections increase dependent upon the increase in the middle class
population. To this end, environmental data from the Organization for Economic
Cooperation and Development is employeda convenient relational source of data, as
Wilson and Dragusanu related the OECD nations as the upper bound of the global
middle class (2). The OECD Environmental Data Compendium Report of 2006-2008
provides un-recycled municipal solid waste data for the participating nations from 1980
to 2005 in two useful forms: on a total (aggregate) basis, and on a per capita basis. The
data shows for the period that the rate of waste generation was large in the 1990s, but
tapered significantlyand slightly declinedfrom the year 2000 to the year 2005, as
depicted in Figure 7-7. Understanding that most of the participating countries are co nsidered developed nations with middle-class majorities provides some insight to this
trend.

74

Figure 7-7: OECD waste generation in kg/capita/year for the period between 1980 and 2005. As all
OECD countries are represented in the legend, only several are listed. (9)

The United States Environmental Protection Agency documents that efforts in


awareness, governmental regulation, and improved recycling solutions have resulted in a
decline in the per capita waste generation rate (11). This point is further illustrated in
Figure 7-8, which may indicate that waste generation rates may have peaked at 4.65
pounds of waste per person per day. The timing of this occurring also corresponds to the
OECD countries waste generation rates as described in Figure 7-7. The U.S. EPA data
also gives insight as to the reason for this decrease in waste generation, as depicted in
Figure 7-9, where recycling is shown to have achieved rates of 33% of the generated
waste stream. Figure 7-10 shows that the relation amongst the OECD and U.S. EPA
findings is more than anecdotal. This relation demonstrates, though the magnitude of the
generated waste estimate is underestimated by the OECD, the general trend of increasing
waste production rate to the eventual leveling of the rate of production from 1980 to
2005 is essentially the same. The difference in the data being a scalar is within 1%,
showing high correlation despite the offset 41.

41

The offset may be due to a variety of factors as mentioned in the assumptions of waste
generation rates and differences in definitions (11) (9).
75

Figure 7-8: U.S. EPA data showing municipal solid waste generation of the United States
between 1960 and 2008 (in million tons). (12)

Figure 7-9: U.S. EPA data showing municipal solid waste recycling (in million tons). (12)

It is the leveling off of the waste generation rate that forms the basis of an important
assumption in the sustainability analysis for waste production. Based on the trends
noted in both the OECD data and the U.S. EPA data, which the amount of waste produced by predominantly middle class developed nations is curbed by recycling and
government enforced regulation, the assumption is formed that the production of waste
76

Figure 7-10: United States waste generation as determined independently by the United States
EPA (11)and the OECD (9).

in a developed nation is asymptotic to a constant. Otherwise stated, there may exist a


natural limit to the amount of waste generated per middle class person per day. The
OECD rate of waste generation from 2004 to 2010 is, conservatively estimated 42, 3.38
lbs of MSW per person per day.
The rate of waste generated is assumed constant for the middle class43, and the projected amount of waste generation tracks linearly to the growth of the global middle
class, as depicted in Figure 7-11. This leads to the staggering number of 2.2 billion tons
of waste generated per annum by the year 2030. Using Tchobanoglous relationship of
roughly 500 lbs. of municipal solid waste per cu bic yard 44 (26), this gives 8.9 billion

42

Assuming that the United States represents the highest extreme of the global middle
class
43
The waste generation rates are assumed to be constant due to factors that contribute to
the leveling seen in the waste generation data. It is noted that regulation in the deve loped countries has helped curb discarded waste and helped establish recycling centers
and programs. If a developing nation were not to institute such regulation or otherwise
encourage conservation and recycling, the rate of waste generation of the global middle
class may mimic the waste generation rates of the developed nations in the years prior to
2000.
44
Estimate based upon the average mass to volume relationship for as -received waste at
a landfill from a waste compactor truck.
77

cubic yards of waste. This amount of waste is sufficient to bury all five boroughs of
New York City 45 in nearly 6 meters [20 feet] of MSW.
Again, the question remains as to whether this amount of waste generation is sustainable. The answer on the basis of available land on a worldwide basis is yes.
However, the definitive answer may be largely economic and political, as the most
preferred method of waste management, landfilling, is expensive and, in every locale,
publicly undesirable.

Figure 7-11: Waste generation as a function of the population of the world wide middle class.
This figure is analytically generated based upon data and projections from the U.S. EPA (12), the
OECD (9), and Goldman Sachs (2).See also Figure 1-3.

45

The land area of New York City is estimated at 780 square kilometers [301 square miles] by the NYC statistics page at http://www.nycgo.com/articles/nyc-statistics-page
(referenced April 2014).
78

7.2 Proximate and Ultimate analysis of waste stream


Table 7-2: Elemental contents of a typical municipal solid waste stream (26)Categories highlighted
are represented constituents of the waste stream in this analysis. **Mixed textiles are not found in
Tchobanoglous text, but is averaged here for simplification of the waste stream analysis.
Type of Waste
Food and Food Products
Fats
Food wastes (mixed)
Fruit wastes
Meat wastes

Carbon

Percent by weight (dry basis)


Hydrogen
Oxygen
Nitrogen

Sulfur

Ash

73.0
48.0
48.5
59.6

11.5
6.4
6.2
9.4

14.8
37.6
39.5
24.7

0.4
2.6
1.4
1.2

0.1
0.4
0.2
0.2

0.2
5.0
4.2
4.9

Paper Products
Cardboard
Magazines
Newsprint
Paper (mixed)
Waxed cartons

43.0
32.9
49.1
43.4
59.2

5.9
5.0
6.1
5.8
9.3

44.8
38.6
43.0
44.3
30.1

0.3
0.1
<0.1
0.3
0.1

0.2
0.1
0.2
0.2
0.1

5.0
23.3
1.5
6.0
1.2

Plastics
Plastics (mixed)
Polyethylene
Polystyrene
Polyurethane
Polyvinyl chloride

60.0
85.2
87.1
63.3
45.2

7.2
14.2
8.4
6.3
5.6

22.8
-4.0
17.6
1.6

-<0.1
0.2
6.0
0.1

-<0.1
-<0.1
0.1

10.0
0.4
0.3
4.3
2.0

Textiles, rubber, leather


Textiles
Rubber
Leather
Textiles (mixed) **

48.0
69.7
60.0
59.2

6.4
8.7
8.0
7.7

40.0
-11.6
25.8

2.2
-10.0
6.1

0.2
1.6
0.4
0.7

3.2
20.0
10.0
11.1

Wood, trees, etc


Yard wastes
Wood (green timber)
Hardwood
Wood (mixed)
Wood chips (mixed)

46.0
50.1
49.6
49.5
48.1

6.0
6.4
6.1
6.0
5.8

38.0
42.3
43.2
42.7
45.5

3.4
0.1
0.1
0.2
0.1

0.3
0.1
<0.1
<0.1
<0.1

6.3
1.0
0.9
1.5
0.4

Glass, metals, etc.


Glass and mineral
Metals (mixed)

0.5
4.5

0.1
0.6

0.4
4.3

<0.1
<0.1

---

98.9
90.5

Miscellaneous
Office sweepings
Oils, paints

24.3
66.9

3.0
9.6

4.0
5.2

0.5
2.0

0.2
--

68.0
16.3

79

7.3 Plasma gasification program


The program used to predict the plasma gasification reaction (including tailoring the
waste stream and estimating the synthesis gas output) is documented here. The code is
written for Mathematica Version 6.0.1. This program is adapted from the program
developed by A. Mountouris denoted as GasifEq, provided courtesy of A. Mountouris
for the purpose of this thesis as a Microsoft Excel spreadsheet with an optimization
macro. The program is inline commented for documentation; however the full methodology of the program and differences from A. Mountouris work is defined in Section 3
of this document.
Notes reading Mathematica syntax:
The programming methodology is procedural, with each module existing ind ividually and called by other modules as required (typically within the main
program)
(* *) - encloses comment - non-executable
Italicized font Indicates section heading non-executable
Boxed Font Module headings non-executable
/ - Indicates line break. Note due to word wrap, Mathematica line breaks may
not coincide with the carriage returns of this document
[[ ]] All coding is matrix and table based. Brackets represent row/column of
data in array
[ ] Functions, modules, and routines are all enclosed in brackets. The full
closure of the brackets indicates the end of the function, module, or routine.
XXXX Values that are readily manipulated as variables in the analysis

80

Plasma Gasification
Optimization Program for Community/Local Waste Consumption and Energy Production
Purpose:
This module is developed from the synthesis gas model of waste gasification
developed by Dr. A. Mountouris.
Methodology:
Optimization electrical output for the case of zero net thermal energy. Based
upon work by Mountouris indicating that the case of zero net thermal energy
results in higher volume of produced synthesis gas. (Sewage
sludge,Mountouris, 2008)

Waste Stream Generation (Ultimate Analysis)


This routine creates the waste stream from the input trash stream. The input
trash stream is post-recycling content of waste with the content of each in
kg/day. The feed rate is in kg/day. The output waste stream is the content
of elements in the waste stream as a percentage of that wastestream.
Note, the sulfur content is so small in the evaluation that it is only utilized for the calculation of the mass
trash[trashStream_, wasteStream_] :=
Module[{organics = trashStream[[1, 1]],
paper = trashStream[[1, 2]],
plastics = trashStream[[1, 3]],
textiles = trashStream[[1, 4]],
wood = trashStream[[1, 5]],
glass = trashStream[[1, 6]],
metals = trashStream[[1, 7]]},
feedrate = (organics) + (paper) + (plastics) + (textiles) +
(wood) + (glass) \
+ (metals);
carbon = ((organics*.480) + (paper*.434) + (plastics*. 600) +
(textiles*.592) \
+ (wood*.481) + (glass*.005) + (metals*.045))/feedrate;
hydrogen = ((organics*.064) + (paper*.058) + (plastics*.072) + \
(textiles*.077) + (wood*.058) + (glass*.001) + (metals*.006))/feedrate;
oxygen = ((organics*.376) + (paper*.443) + (plastics*.228) +
(textiles*.258) \
+ (wood*.455) + (glass*.004) + (metals*.043))/feedrate;
nitrogen = ((organics*.026) + (paper*.003) + (plastics*.000) + \
(textiles*.061) + (wood*.001) + (glass*.000) + (metals*.000))/feedrate;
sulfur = ((organics*.004) + (paper*.002) + (plastics*.000) +
(textiles*.007) \
+ (wood*.001) + (glass*.000) + (metals*.000))/feedrate;
ash =
((organics*.05) + (paper*.006) + (plastics*.100) +
(textiles*.111) \

81

+ (wood*.004) + (glass*.989) + (metals*.905))/feedrate;


moisture = ((organics*.600) + (paper*.060) + (plastics*.020) + \
(textiles*.060) + (wood*.350) + (glass*.000) + (metals*.000))/feedrate;
(* Waste Stream Input Order
Carbon (%),
Hydrogen(%),
Nitrogen(%),
Oxygen(%),
Ash(%),
Moisture(%),
feedrate (kg/d)
*)
wasteStream = {{carbon*100, hydrogen*100, nitrogen*100, oxygen*100,
ash*100,
moisture*100, feedrate}};
]

82

Waste Stream (Molar Fraction Basis)


This routine provides the molecular basis of the input waste stream
Variable List:
molar fractions -- nn=Nitrogen, oo=oxygen, hh=hydrogen, cc=carbon
molar basis of waste wasteMB
waste[wasteStream_, wasteMB_, molarFraction_] :=
Module[{ c = wasteStream[[1, 1]], h = wasteStream[[1, 2]],
o = wasteStream[[1, 4]], n = wasteStream[[1, 3]],
nn = 0, cc = 0, hh = 0, oo = 0},
nn
oo
hh
cc

=
=
=
=

(n/14)/(c/12);
(o/16)/(c/12);
(h/1)/(c/12);
(c/12)/(c/12);

wasteMB = cc*12 + oo*16 + hh*1 + nn*14;


molarFraction = {{cc, hh, nn, oo}};
]

83

Waste Stream Raw Heat Capacity


Heat Capacity
The following data is from the Chemical Properties Handbook.
copied from Mountouris, but confirmed in the CPH.

The values are

heatCapacity[heatCP_] :=
Module[{heatH2 = 0, heatN2 = 0, heatCO = 0, heatCO2 = 0, heatC = 0,
heatO2 = 0, heatCH4 = 0, heatH2Og = 0},

heatH2 = {2.5399*10^+1, 2.0178*10^ -2, -3.8549*10^-5,


3.1880*10^-8, -8.7585*10^-12};
heatO2 = {2.9526*10^+1, -8.8999*10^-3, 3.8083*10^-5, -3.2629*10^-8,
8.8607*10^-12};
heatN2 = {2.9342*10^+1, -3.5395*10^-3, 1.0076*10^-5, -4.3116*10^-9,
2.5935*10^-13};
heatCO = {2.9556*10^+1, -6.5807*10^-3, 2.0130*10^-5, -1.2227*10^-8,
2.2617*10^-12};
heatCO2 = {2.7437*10^+1, 4.2315*10^-2, -1.9555*10^-5,
3.9968*10^-9, -2.9872*10^-13};
heatCH4 = {3.4942*10^+1, -3.9957*10^-2, 1.9184*10^-4, -1.5303*10^-7,
3.9321*10^-11};
heatH2Og = {3.3933*10^+1, -8.4186*10^-3, 2.9906*10^-5, -1.7825*10^-8,
3.6934*10^-12};
heatC = {-8.3200*10^-1, 3.4846*10^-2, -1.3233*10^-5, 0, 0};
heatCP = Table[0, {8}, {5}];
heatCP[[1]] = heatC;
heatCP[[2]] = heatCH4;
heatCP[[3]] = heatCO;
heatCP[[4]] = heatCO2;
heatCP[[5]] = heatH2;
heatCP[[6]] = heatH2Og;
heatCP[[7]] = heatO2;
heatCP[[8]] = heatN2;
(*
Debugging
Print["HeatCapacity"];
Print[MatrixForm[heatCP]];*)
]

84

Heat of Formation
The following data is from the Chemical Properties Handbook
The Boie relation is employed to determine the HHV of the waste stream and
subsequently the heat capacity of the waste (as an input fuel)
heatFormation[wasteMB_, wasteStream_, heatForm_] :=
Module[{wMB = wasteMB, carbon = wasteStream[[1, 1]],
hydrogen = wasteStream[[1, 2]], nitrogen = wasteStream[[1, 3]],
oxygen = wasteStream[[1, 4]]},
heatForm = Table[0, {1}, {10}];
heatForm[[1,
heatForm[[1,
H2Ol *)
heatForm[[1,
heatForm[[1,
heatForm[[1,
heatForm[[1,

2]] = -241800; (* H2Og *)


3]] = -285830.4; (*
4]]
5]]
6]]
7]]

=
=
=
=

0.; (* O2 *)
0.;
0.; (* H2 *)
-110500; (* CO *)

heatForm[[1, 8]] = -393500; (* CO2 *)


heatForm[[1, 9]] = -74850; (* CH4 *)
heatForm[[1, 10]] =
40806 (* Heat of vaporization at 373K (the boiling point of water)*)
ClearAll[boieHHV];
boieHHV = (35160*carbon/100) + (116225*hydrogen/100) - (11090*
oxygen/100) + (6280*
nitrogen/
100); (*Calculate HHV of waste stream based on Boie Relationship *)
heatForm[[1, 1]] = (wMB*boieHHV) + heatForm[[1, 8]] +
heatForm[[1, 3]]*(((.01*hydrogen/1.01)/(.01*carbon/12))/2);
(*
Print["boieHHV: ",boieHHV];
Print["Waste Heat of Formation: ",heatForm[[1,1]]];
Print["dafMB:",wMB];
Print[];
*)
]

85

Waste Mass Basis


massE[wasteStream_, wasteMB_, daf_, dafMB_, ashAW_, ashMass_] :=
Module[{wmb = wasteMB, a = wasteStream[[1, 5]],
f = wasteStream[[1, 7]]/(24*60*60), mc = wasteStream[[1, 6]]},
daf = f*(1 - (mc/100))*(1 - (a/100)); (* kmol/s *)
dafMB = (f/wmb)*(1 - (mc/100))*(1 - (a/100)); (* kg/s *)
ashMass = (daf*100/(100 - a)) - daf;
ashAW = (100/f)*ashMass
(*
Print["daf:",daf];
Print["dafMB:",dafMB];
Print["ashMass:", ashMass];
Print["ashAW:",ashAW]
*)
]

86

Equilibrium Balance
This routine performs the chemical equilibrium balance and is the h eart of
the GasifEq system
eQuil[wasteStream_, wasteMB_, ashAW_, molarFraction_, heatForm_, heatCP_,
massO2_, dafMB_, tempSynG_, tempAmb_, tempDry_, solution_] :=
Module[{mc = 0, water = 0, nEquil = 0, balC = 0, balH = 0, balO = 0,
heatBalance = 0,
x, y, z, nrg, cpMethane, cpWaterg, gibbs, k1 = 0, k2 = 0,
lnklnk0 = 0,
lnk0 = 0, lnk = 0},
Clear[nrg, x, y, z]; (* nrg = electricity, x = nH2, y= nCO, z=nCO2 *)
Clear[cpMethane, cpWaterg, gibbs];
(* Gibbs Free Energy in this order: H2 O2 N2 CO CO2 CH4 H2O(g) H2O(l)*)
gibbs = {0, 0, 0, -137.28, -394.38, -50.84, -228.6};
(*Equilibrium Reactions*)
(* CH4 + H2O = CO +
3H2 -- Methane Decomposition*)
cpMethane =
Table[heatCP[[3, i]] + 3*heatCP[[5, i]] - heatCP[[2, i]] heatCP[[6, i]], {i, 1, 5}];
(*Print["cpMethane: ",cpMethane];*)
lnklnk0 = ((cpMethane[[1]]/8.314)*
Log[
tempSynG/298.15]) + ((cpMethane[[2]]/(2*8.314))*(tempSynG - 298.15)) +
((cpMethane[[3]]/(6*8.314))*((tempSynG ^2) - (298.15^2))) +
((cpMethane[[4]]/(12*8.314))*((tempSynG^3) - (298.15^3))) +
((cpMethane[[5]]/(20*8.314))*((tempSynG^4) - (298.15^4))) ((((heatForm[[1, 7]] + 3*heatForm[[1, 6]] - heatForm[[1, 9]] heatForm[[1, 2]])/(8.314)) (1/
8.314)*(cpMethane[[
1]]*298.15 + (cpMethane[[2]]/2)*(298.15^2) + (cpMethane[[3]]/
3)*(298.15^3) +
(cpMethane[[4]]/4)*(298.15^4) + (cpMethane[[5]]/
5)*(298.15^5)))*((1/tempSynG) - (1/298.15)));
lnk0 = -(gibbs[[4]] + 3*gibbs[[1]] - gibbs[[6]] gibbs[[7]])/(8.314*298.15/
1000);
k1 = Exp[lnklnk0 + lnk0];
(* CH4 + H2O = CO + 3H2 -- Water Gas Shift*)
cpWaterg =
Table[heatCP[[4, i]] + heatCP[[5, i]] - heatCP[[3, i]] -

87

heatCP[[6, i]], {i, 1, 5}];


(*Print["cpWaterg: ",cpWaterg];*)
lnklnk0 = ((cpWaterg[[1]]/8.314)*
Log[tempSynG/298.15]) + ((cpWaterg[[2]]/(2*8.314))*(tempSynG - 298.15))
+
((cpWaterg[[3]]/(6*8.314))*((tempSynG^2) - (298.15^2))) +
((cpWaterg[[4]]/(12*8.314))*((tempSynG^3) - (298.15^3))) +
((cpWaterg[[5]]/(20*8.314))*((tempSynG^4) - (298.15^4))) ((((heatForm[[1, 8]] + heatForm[[1, 6]] - heatForm[[1, 7]] heatForm[[1, 2]])/(8.314)) (1/8.314)*(cpWaterg[[1]]*298.15 +
(cpWaterg[[2]]/2)*(298.15^2) +
(cpWaterg[[3]]/3)*(298.15^3) +
(cpWaterg[[4]]/4)*(298.15^4) +
(cpWaterg[[5]]/5)*(298.15^5)))*((1/tempSynG) (1/298.15)));
lnk0 = -(gibbs[[5]] + gibbs[[1]] - gibbs[[4]] - gibbs[[7]])/(8.314*298.15/
1000);
k2 = Exp[lnklnk0 + lnk0];
(*Set up the reaction coefficients in the equilibrium reaction*)
mc = (100/(100 - ashAW))*wasteStream[[1, 6]];
water = wasteMB*mc/(18*(100 - mc));
massO2 = .47;(*nO2=assumed value -- when working replace with a slider*)
nEquil = Table[0, {1}, {6}];
nEquil[[1, 6]] = (massO2*3.76) + (molarFraction[[1, 3]]/2);(*nN2*)
nEquil[[1, 1]] = x; (*nH2=Initial guess*)
nEquil[[1, 2]] = y; (*nCO=
Initial guess*)
nEquil[[1, 3]] = z; (*nCO2=Initial guess*)
nEquil[[1, 4]] = nEquil[[1, 3]]*nEquil[[1, 1]]/(k2*nEquil[[1, 2]]);(*nH2O*)
nEquil[[1,
5]] = (nEquil[[1, 1]]^3)*(nEquil[[1,
2]]/(nEquil[[1, 4]]*k1*(nEquil[[1, 1]] +
nEquil[[1, 2]] + nEquil[[1, 3]] + nEquil[[1, 4]] +
nEquil[[1, 6]])^2));(*nCH4*)
(*Print["nEquil:",nEquil];*)
(* Equations for Balances *)
balC = nEquil[[1, 2]] + nEquil[[1, 3]] + nEquil[[1, 5]] molarFraction[[1, 1]];
balH = nEquil[[1, 1]] + nEquil[[1, 4]] + (2*nEquil[[1, 5]]) water - (molarFraction[[1, 2]]/2);
balO = nEquil[[1, 2]] + (2*nEquil[[1, 3]]) + nEquil[[1, 4]] - 2*massO2 water - molarFraction[[1, 4]];

88

heatBalance = heatForm[[1, 1]] + ((water)*(heatForm[[1, 3]])) +


massO2*heatForm[[1, 4]] + nEquil[[1, 6]]*heatForm[[1, 5]] +
(nrg/dafMB) - nEquil[[1, 1]]*heatForm[[1, 6]] nEquil[[1, 2]]*heatForm[[1, 7]] - nEquil[[1, 3]]*heatForm[[1, 8]] nEquil[[1, 4]]*heatForm[[1, 2]] - nEquil[[1, 5]]*heatForm[[1, 9]] (nEquil[[1,
1]]*(heatCP[[5,
1]]*(tempSynG - tempAmb) + (heatCP[[5, 2]]/2)*(tempSynG^2 tempAmb^2) +
(heatCP[[5, 3]]/3)*(tempSynG^3 - tempAmb^3) + (heatCP[[5,
4]]/
4)*(tempSynG^4 - tempAmb^4) +
(heatCP[[5, 5]]/5)*(tempSynG^5 - tempAmb^5)) +
nEquil[[1,
2]]*(heatCP[[3,
1]]*(tempSynG - tempAmb) + (heatCP[[3, 2]]/2)*(tempSynG^2 tempAmb^2) +
(heatCP[[3, 3]]/3)*(tempSynG^3 - tempAmb^3) + (heatCP[[3,
4]]/
4)*(tempSynG^4 - tempAmb^4) +
(heatCP[[3, 5]]/5)*(tempSynG^5 - tempAmb^5)) +
nEquil[[1, 3]]*(heatCP[[4, 1]]*(tempSynG - tempAmb) +
(heatCP[[4, 2]]/2)*(tempSynG^2 - tempAmb^2) + (heatCP[[4,
3]]/
3)*(tempSynG^3 - tempAmb^3) +
(heatCP[[4, 4]]/4)*(tempSynG^4 - tempAmb^4) + (heatCP[[4,
5]]/
5)*(tempSynG^5 - tempAmb^5)) +
nEquil[[1,
4]]*(heatCP[[6,
1]]*(tempSynG - tempAmb) + (heatCP[[6, 2]]/2)*(tempSynG^2 tempAmb^2) +
(heatCP[[6, 3]]/3)*(tempSynG^3 - tempAmb^3) + (heatCP[[6,
4]]/
4)*(tempSynG^4 - tempAmb^4) +
(heatCP[[6, 5]]/5)*(tempSynG^5 - tempAmb^5)) +
nEquil[[1,
5]]*(heatCP[[2,
1]]*(tempSynG - tempAmb) + (heatCP[[2, 2]]/2)*(tempSynG^2 tempAmb^2) +
(heatCP[[2, 3]]/3)*(tempSynG^3 - tempAmb^3) + (heatCP[[2,
4]]/
4)*(tempSynG^4 - tempAmb^4) +
(heatCP[[2, 5]]/5)*(tempSynG^5 - tempAmb^5)) +
nEquil[[1,
6]]*(heatCP[[8,
1]]*(tempSynG - tempAmb) + (heatCP[[8, 2]]/2)*(tempSynG^2 tempAmb^2) +
(heatCP[[8, 3]]/3)*(tempSynG^3 - tempAmb^3) + (heatCP[[8,
4]]/
4)*(tempSynG^4 - tempAmb^4) +
(heatCP[[8, 5]]/5)*(tempSynG^5 - tempAmb^5)));
(*

89

Print["Carbon Balance: ", balC];


Print["Hydrogen Balance: ", balH];
Print["Oxygen Balance: ", balO];
Print["Heat Balance: ", heatBalance];
Print["HeatCP"];
Print[MatrixForm[heatCP]];
Print["Equilibrium (H2, CO, CO2, H2O, CH4, N2): ",MatrixForm[nEquil]];
*)
solution =
NSolve[{heatBalance == 0, balO == 0, balH == 0, balC == 0}, {x, y, z,
nrg}];
x = x /. solution;
y = y /. solution;
z = z /. solution;
nrg = nrg /. solution;
(*
Print["Equilibrium (H2, CO, CO2, H2O, CH4, N2): "];
Print[MatrixForm[nEquil]];
*)
(*Print[x];Print[y];Print[z];Print[nrg];*)
ClearAll[solution]; solution = Table[0, {7}];
Do[
If[x[[i]] > 0 && y[[i]] > 0 && z[[i]] > 0,
solution[[1]] = x[[i]];
solution[[2]] = y[[i]];
solution[[3]] = z[[i]];
solution[[7]] = nrg[[i]]
]
, {i, Length[nrg]}
]
nEquil[[1, 4]];
solution[[4]] = Last[nEquil[[1, 4]]];
solution[[5]] = Last[nEquil[[1, 5]]];
solution[[6]] = nEquil[[1, 6]]
(*Print[solution];*)
(*

Print[NMinimize[{heatBalance,
{nrg,{x,0,3},y,z}]];
*)
(*
Print["Carbon Balance: ", balC];
Print["Hydrogen Balance: ", balH];
Print["Oxygen Balance: ", balO];

90

Print["Heat Balance: ", heatBalance];


Print["x = nH2, y= nCO, z=nCO2, nrg"];
solution={nrg,x,y,z}/.%;
*)
]

91

Main Program
1) Recycling (waste stream conditioning)
2) Proximate Analysis (breakdown of waste stream by waste type)
3) Executes all previous modules to determine synthesis gas production and
energy consumption
4) Outputs data to screen or Excel for additional results manipulation

(*Define Waste Stream*)


(* This function is only temporary as a check of the \
plasma module.
In later sections, this code will actually reside in the waste stream module.
\
*)

(* This is the waste stream objective function --- the order of constituents
\
are:
Carbon,
Hydrogen,
Nitrogen,
Oxygen in percent of waste stream and
Ash,
Moisture (added to 100%) and
feedrate (kg/d)
*)

(*Clear[counter,results];
Clear[tempSynGHigh,tempSynGLow, tempStep];

tempSynGHigh=1500;tempSynGLow=1500;tempStep=100;
moistHigh=80;moistLow=20;moistStep=10;
solutionSize= \
(((moistHigh-moistLow)/moistStep)+1)*(((tempSynGHightempSynGLow)/tempStep)+1);
counter=0; results=Table[0,{solutionSize},{3}];
*)
(*Literature indicates plasma gasification temperatures of 2200F.
This is outside of the range of the equations selected for the study from the
Chemical Properties Handbook. Such a high temperature ensures the absence of
solide carbon(soot) *)
arraySize = 10000;
iterates,

(* This value sets the number o f times the program \


each time modifying the inputs to the wast-

estream *)

92

ClearAll[solutionSet];
ClearAll[inputSet];
ClearAll[wasteinput];
solutionSet = Table[0, {arraySize}];
inputSet = Table[0, {arraySize}];
tempAmb =
273 + 10.7; (* Ambient temperature based on average daily New England
temperature for full year operation (K ) *)
tempDry = 373; (* Drying temperature - this will also later be a variable
(K)*)
wasteinput = 10; (* kg/day*)
tempSynG = 1500; (* Temperature of the syntheis gas (K) *)

Do[
(* wasteStream= Carbon, Hydrogen, Nitrogen, Oxygen, Ash *)
(* m = moisture \
content, fr = Feed Rate, t = temperature *)

ClearAll[heatCP, heatForm];
ClearAll[wasteMB, dafMB , daf, ashAW, ashMass, molarFraction, massO2];
ClearAll[solution];
ClearAll[wasteStream];
ClearAll[trashStream];
ClearAll["*'Recycling"];

(* Set initial Values *)


(* Recycling inputs -- 0 = completely recycled \
(not part of the waste stream)
1 = no recycling (full component of the
waste stream)
Assume that recycling will not be
perfect and 5% will still
make it into the plasma unit
Justify with recycling pollution
rates*)
(*
These are test values \
for recycling for spot checks of the optimization or various part s of the \
program
organicsRecycling= 1;
paperRecycling= 1;
plasticsRecycling= 1;
textilesRecycling= 1;
woodRecycling= 1;
glassRecycling= 0;
metalsRecycling= 0;

93

*)
(*
These are the values used for the optimization to allow for the ran domly \
generated waste streams
*)
organicsRecycling =
RandomReal[
1]; (* If[organicsRecycling<.5, organicsRecycling=.05,organicsRecycling=1]; *)
paperRecycling =
RandomReal[
1]; (* If[paperRecycling<.5, paperRecycling=.05 ,paperRecycling=1]; *)
plasticsRecycling =
RandomReal[
1]; (* If[plasticsRecycling<.5, \
plasticsRecycling=.05,plasticsRecycling=1]; *)
textilesRecycling =
RandomReal[
1]; (* If[textilesRecycling<.5, \
textilesRecycling=.05,textilesRecycling=1]; *)
woodRecycling =
RandomReal[
1]; (* If[woodRecycling<.5, woodRecycling=.05,woodRecycling=1]; *)
glassRecycling =
RandomReal[
1]; (* If[glassRecycling<.5, glassRecycling=.05,glassRecycling=1]; *)
metalsRecycling =
RandomReal[
1]; (* If[metalsRecycling<.5, metalsRecycling=.05,metalsRecycling=1]; *)

(* The trash Stream is from Table 1 of the Waste Generation Rates


Rubber & Leather are combined with textiles - assumed similar
content as \
noted on Sheet 1
Inorganic wastes are added as well to the glass - assumed similar content \
as glass *)
trashStream = {{
(.26*wasteinput*organicsRecycling),
(.31*wasteinput*paperRecycling),
(.12*wasteinput*plasticsRecycling),
(.08*wasteinput*textilesRecycling),
(.07*wasteinput*woodRecycling),
(.08*wasteinput*glassRecycling),
(.08*wasteinput*metalsRecycling)

94

}};
trash[trashStream,
wasteStream];

(* Call waste function *)

waste[wasteStream, wasteMB,
molarFraction];
(* Call waste function *)
heatCapacity[
heatCP];
\

(* Call heat capacity function *)

massE[wasteStream, wasteMB, daf, dafMB, ashAW,


ashMass];
(* call mass equation *)
heatFormation[wasteMB, wasteStream,
heatForm];
*)

(* Call heat capacity function

eQuil[wasteStream, wasteMB, ashAW, molarFraction, heatForm,


heatCP, massO2, dafMB, tempSynG, tempAmb, tempDry,
solution]; (* Run Equilibrium Module *)
inputSet[[i]] = {tempSynG, organicsRecycling, paperRecycling,
plasticsRecycling,
textilesRecycling, woodRecycling, glassRec ycling, metalsRecycling};
solutionSet[[i]] = solution
, {i, 1, arraySize}
]
(* This print routine is to check the output file visually instead of \
writing it to a file interpretable by MSExcel
Print[TableForm[Insert[inputSet,{"PlasTemp","Organics","Paper","Plastic",\
"Textiles","Wood","Glass","Metals"},1]]];
Print[TableForm[Insert[wasteStream,{"carbon \
%","hydrogen%","nitrogen%","oxygen%","ash%","moisture%","feedrate"},1]]]; \
Print[];
Print[TableForm[Insert[solutionSet,{"nH2", "nCO", "nCO2", "nH2O", "nCH4", \
"nN2", "Energy Input"},1]]];
*)
(* This print routine writes results to an output file interpretable by \
MSExcel *)
Export["1500_inputSet_large.xls", TableForm[Insert[inputSet,
{"PlasTemp", "Organics", "Paper", "Plastic", "Textiles", "Wood",
"Glass",
"Metals"}, 1]], "XLS"]; Print[];
Export["1500_solutionSet_large.xls", TableForm[Insert[solutionSet,
{"nH2", "nCO", "nCO2", "nH2O", "nCH4", "nN2", "Energy Input"},
1]],
"XLS"];

95

(* This print routine was also initially used to observe the inputs as the \
program operated
Print["Trash Stream"];
Print[trashStream[[1,1]],"
Print[trashStream[[1,2]],"
Print[trashStream[[1,3]],"
Print[trashStream[[1,4]],"
Print[trashStream[[1,5]],"
Print[trashStream[[1,6]],"
Print[trashStream[[1,7]],"
Print[""];

Organics (kg/day)"];
Paper (kg/day)"];
Plastics (kg/day)"];
Textiles (kg/day)"];
Wood (kg/day)"];
Glass (kg/day)"];
Metals (kg/day)"];

Print["Waste Stream"];
Print[wasteStream[[1,1]],"
Print[wasteStream[[1,2]],"
Print[wasteStream[[1,3]],"
Print[wasteStream[[1,4]],"
Print[wasteStream[[1,5]],"
Print[wasteStream[[1,6]],"
Print[wasteStream[[1,7]],"
Print[""];

Carbon (%)"];
Hydrogen (%)"];
Nitrogen (%)"];
Oxygen (%)"];
Ash (%)"];
Moisture (%)"];
Feed rate (kg/day)"];

Print[MatrixForm[solution]];
Print["nH2, nCO, nCO2, nH2O, nCH4, nN2, Energy Input"];
Print[""];
*)
"END"

96

7.4 Chemical properties


7.4.1 Heat Capacities
The following table of heat capacities was compiled from the Chemical Properties
Handbook (22)
Table 7-3: Heat capacities of elements and compounds of gasification reaction model as compiled from
the Chemical Properties Handbook (22)
H2

O2

N2

CO

T regime

250-1500K

50-1500K

50-1500K

60-1500K

2.5399E+01

2.9526E+01

2.9342E+01

2.9556E+01

2.0178E-02

-8.8999E-03

-3.5395E-03

-6.5807E-03

-3.8549E05
3.1880E-08
-8.7585E12

3.8083E-05

1.0076E-05

2.0130E-05

-3.2629E-08
8.8607E-12

-4.3116E-09
2.5935E-13

-1.2227E-08
2.2617E-12

D
E

T regime

CO2

CH4

H2O (g)

50-5000K

50-1500K

250-1500K

200-1100K

2.7437E+01

3.4942E+01

3.3933E+01

-8.3200E-01

4.2315E-02

-3.9957E-02

-8.4186E-03

3.4846E-02

-1.9555E05
3.9968E-09
-2.9872E13

1.9184E-04

2.9906E-05

-1.3233E-05

-1.5303E-07
3.9321E-11

-1.7825E-08
3.6934E-12

D
E

7.4.2 Gibbs free energy


The following table of Gibbs free energy was compiled from the Chemical Prope rties Handbook (22)
Table 7-4: Standard Gibbs Free Energy of Molecular Compounds Presented in the Plasma Gasification
Reaction (22)
CO

CO2

CH4

H2O (g)

T regime

60-1500K

50-5000K

50-1500K

250-1500K

Gf,298.15 (KJ/mol)

-137.28

-394.38

-50.84

-228.6

97

7.4.3 Heats of formation


The following table of heats of formation was compiled from the Chemical Properties Handbook (22).
Table 7-5: Standard Heats of formation of molecular compounds presented in the plasma gasification
reaction (22)
CO

CO2

CH4

H2O (g)

H2O (l)

T regime

60-1500

50-5000

50-1500

250-1500

250-1500

f,298.15
(KJ/mol)

-110.5

-393.5

-74.85

-241.800

-285.8304

98

Hv apor
44.0304

7.5 Single-stream recycling contamination


Contamination of single stream recycled waste content is a real concern. Single stream recycling collection is implemented primarily as a means of reducing the co mplexity of recycling while boosting the amount of material that can be recovered.
Though it is not the intent of this section to argue for or against the implementation of
single-stream recovery efforts, it is notable that such efforts promote the recovery of
materials more than multi-stream methods. The most likely reason for this is the ease of
consumer use.
Making up for the consumer, various innovative and complex recovery systems
have been put in place to maximize materials recovery. Employing segregation methods
ranging from manual separation (at the most basic level) to computerized vision dete ction (among the most complex of these systems) still results in recycled content rejection
due to contamination. The ultimate end-state of this contaminated recycling content is
all too often the municipal solid waste landfill system.46 Studies performed indicate
recycled content losses for paper to be 16%, 7% for aluminum, 32% for plastics (54).
With the introduction of plasma waste reforming, the diligent efforts for waste recovery may no longer be necessary. Most of the recovered content would go into the
production of nitrogen, carbon dioxide, water, and electrical power. The recovered
inorganic solids are removed from the plasma reactor crucible as metals and a vitreous,
glassy material, called slag. The slag has potential for a variety of building and co nstruction materials, and the aggregated metals can be metallurgically separated 47 to
return them to more useful forms.
Because of this, the analysis presented here takes into account recycled stream co ntamination on levels of 5-10% of the waste stream overall. Though power requirements
and hydrogen production are not optimized with the introduction of these materials int o
the waste stream, the presence of the contamination is important enough to plan for in
any commercially viable solution. The influence of this waste stream contamination is

46

At least in the United States.


By tapping the crucible of its molten contents at different levels, solid metals and
vitrified solids may be separated.
99
47

significant, but do not provide enough detriment to negate the feasibility of a plasma
waste gasification system on an operational or economical basis.

7.6 Source reduction


It is important to address source reduction in the discussion regarding any waste-toenergy method, such as plasma gasification. The conversion of waste into energ y using
plasma gasification technology requires waste as an input; however this should not be
seen to conflict with efforts to reduce the sources of waste. Instead, the technology
should be seen as a synergistic means to an end: as long as waste is generated, its benign
disposal and residual energy recovery are necessary. Never should waste be generated
as a means to create electrical power.
High-energy waste is also desirable for plasma gasification. Plastics and any petroleum or oil based product (e.g. rubber, coated papers, and Styrofoam) and even paper
promote the production of a higher energy synthesis gas. However, directly recycling
these products is also highly desirable as the costs and materials (as well as the energy)
to regenerate products from these forms are high. If plasma gasification of waste is
pursued in mass, it cannot be at the expense of plant matter (paper based) and petroleum
(oil based) products because the recycled material reclaimed in plasma gasification is
energy, not material. Since the basic input materials are not being recycled, source
reduction is of significant importance. Means of reducing the amount of paper and
petroleum based products consumed is still fundamental in a comprehensive sustainability model.

7.7 Conveyor sizing


Sizing of the conveyor is based on a horizontal feed system. The system is sized
based on the amount of waste required to provide sufficient synthesis gas to power the
operation. This is an iterative process, where it was first assumed that the h orsepower
requirements to transport and compress the waste is 1/5 that of what was used to grind it.
That provided for a system that required 0.7 kW based upon sizing the power require-

100

ments from a commercially available electric motor48. Equations for sizing the conveyor
(55) were then employed based upon the waste feed rate requirements to provide sufficient hydrogen to fuel the gasification process. The equation used is (55):

[25]

where power P, is given in kW and PH is the power for progressing the material, PN is
the driving power of the screw conveyor, and PST is the power required due to the
incline of the conveyor. As this conveyor is horizontal, PST is zero. PH is (55)

[26]

in which IM is the mass flow rate of media in metric tons per hour, L is the length of the
conveyor, and is the friction coefficient resulting from the conveyed media. The mass
flow rate of waste is was directly determined from the gasification program, the length
was determined by sizing the diameter of the conveyor to be 0.6 m [2 ft], which is
reasonably sized based on dimensions of the waste shredder (50). Knowing that municipal solid waste when compacted occupies approximately 297 kg/m3 [500 lb/yd 3] (10)and
the waste introduction rated based upon the results of the gasification system for selfsufficient power is 0.16 metric tons/hr [320 lb/hr or 4 tons US/day], the length of the
conveyor was determined to be 1.68 m [5.5 ft]49. The friction term typically ranges
between 2 and 4 (55), but the conveyed media in the lower portion of this range is rock.
MSW is expected be to much less than this . Instead

48

is used

. Finally

Sized from a MicroMax motor operating at 230V, 3.2 amp, 1 HP electric motor from
Marathon Electric: www.automationdirect.com/motors . Volume 14. p. E15-32.
49
Note that this is a conveyor feeding from both sides so as to slightly compact the
waste in the middle prior to introduction into the plasma gasification chamber.
101

[27]

where D is the diameter of the conveyor. Proceeding through these calculations given in
equations [25], [26], and [27] shows that the conveyor may consume 1.2 kW of power.

102

Das könnte Ihnen auch gefallen