Undergraduate’s Guide to Writing Chemistry Papers
2018-2019 Edition
Binyomin Abrams
Department of Chemistry, Boston University
Copyright
c
2017, Binyomin Abrams. All rights reserved. This book or parts thereof may
not be reproduced in any form, stored in any retrieval system, or transmitted in any form by
any means – electronic, mechanical, photocopy, recording, or otherwise without prior written
permission of the author, except as provided by United States of America copyright law. For
permission requests, write to the author, at:
Binyomin Abrams
Department of Chemistry, Boston University
Boston, MA 02215, USA
2
Table of Contents
Writing 1: Introduction to Scientific Writing 3
Writing 2: Preparing Appropriate Exhibits 11
Writing 3: Making Claims and Building an Argument 23
Writing 4: Structure of Abridged Scholarly Papers 33
Writing 5: Conventions of Scientific Writing 45
Writing 6: Research, Scientific Literature, and Annotated Bibliographies 61
Writing 7: Adding Motivation Introduction Sections 77
Writing 8: Sharing your Process Experimental Sections 85
Writing 9: Getting your Work Funded Research Proposals 95
Writing 10: Making and Delivering Effective Research Presentations 99
2 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 1: INTRODUCTION TO SCIENTIFIC WRITING
Before we can begin discussing how to approach the scientific writing that you will do in this
course, we must first preface with a little bit of background about this type of writing technical
writing for the sciences. While scientific writing shares many common aspects with other forms
of argumentative writing, discipline-specific conventions and nuances also play an important
role. Although this guide will focus on the practices important in writing chemistry papers,
most of the details are common across other scientific disciplines.
1.1 Why do we write?
If you ask most scientists, they’d probably prefer to avoid the writing altogether and focus
solely on the actual science. That said, when pressed, they will admit that it is also one of
the most crucial and time-consuming parts of what they do. Communication about scientific
results is absolutely essential to the enterprise of science. Discovery without communication is
pointless.
There are many different types of writing that a career scientist engages in. Grant proposals
are the main vehicle by which scientists petition funding agency for money to maintain their
research programs. A researcher will spend a significant portion of time and effort writing
grants and, if they are successful in acquiring funding, they will have to write periodic progress
reports and summaries to the agencies.
Once the experiments are complete, the final step in the Scientific Method involves the
dissemination of findings through two different, but simultaneous, approaches that a researcher
will take in order to spread his work into the greater scientific community: (1) make oral
presentations at conferences, and (2) publish scholarly papers. Publication is considered the
most desirable and necessary method of communication as it creates an accredited record of
the results and allows for the greatest dissemination of the information across both distance
and time. In general, the purpose of publishing a scholarly paper is to convey the findings of an
experiment(s) that has been performed (primary), to persuade the greater scientific community
to adopt one’s own perspective on a given topic (secondary), and to justify the funding that
has been received (tertiary).
Scientific research is a team sport. After individual research groups perform experiments,
and those experiments are validated and duplicated, they attempt to persuade others to accept
or reject their hypotheses by presenting and interpreting the data. The scholarly paper is the
vehicle of persuasion. After the paper has been submitted, it is given to other scientists for
review. If the findings stand up to criticism, they become part of the accepted body of scientific
3
4 · Undergraduate’s Guide to Writing Chemistry Papers
knowledge until they are later disproved. These scholarly papers make up the primary sources
in scientific literature.
1.2 What do we write?
A research scientist will never write a ‘Lab report.’ While it is unclear as to the exact historical
origin of the dreaded lab report, it is clear that only students in science courses are ever asked
to produce these ridiculous types of papers. The sole goal of the lab report is to convey the
specific findings that a student achieved in a pseudo-scientific situation, where the outcome of
the experiment is well-known and documented, to a teacher looking for the results. Outside of
undergraduate instructional courses, this type of paper simply does not exist. These papers are
absent of any scientific motivation (clarification: a grade is not motivation) and are plagued
with stylistic horrors that are designed to streamline the grading process.
Whatever the history, or reasons, behind the lab report, it is clear that it is not a genuine
form of scientific communication. It is for this reason that we do not believe in teaching that
form of writing. Instead, we will focus on developing the skills necessary to formulate a cogent
and persuasive scientific argument in a manner that is consistent with the writing found in
Journal Articles, the most ubiquitous vehicle of scientific communication in the literature. This
approach is geared toward making students good scientists, rather than just good students.
Some, but certainly not all, of the major differences between a lab report and a journal
article are:
Motivation: in journal articles, the Introduction will typically include a significant amount
of details about the motivation behind the project/experiment. Here, motivation refers to
an over-arching reason why the science, or the specific outcome, has global ramifications
and interest. Why were the researchers interested in pursuing this research? Why were
funding agencies willing to fund the research? And why might other people be interested
in the results? In lab reports, students are not motivated by external factors; rather,
they are performing a ‘cookie-cutter’ experiment that is mandated by faculty for the sole
purpose of getting a grade.
Methods: the experimental protocol described in a journal article is a novel approach,
or variation on an approach, that is being published. For students writing lab reports,
however, the formulaic laboratory procedure is not novel and usually does not need to
be presented in such detail. Certainly, routine tasks like cleaning glassware or titrations
would never be described in the scientific literature.
Sample calculations: many instructors prefer to see a student’s sample calculations and
raw data presented in the Results section of lab reports. Neither calculations nor raw
data are ever presented in the body of a true scientific paper or journal article. Rather,
the important formulas are presented along with the final results, and the rest of the
information (data and calculations) will be presented in the rare occasions that they
are sophisticated enough or interesting enough to be included at all in the Supporting
Information section (more on this later).
Arguments vs. data dump: the writing in journal articles must be extremely persuasive,
as the entire function is to promote a set of findings to the broader scientific community.
Writing 1: Introduction to Scientific Writing · 5
When discussing the validity and consequence of these results, it is critical that the ar-
gument presented be compelling. Lab reports, however, are usually seen as a vehicle for
the dissemination of a student’s results to an instructor. As such, it is rarely necessary
for the student to justify results, rather simply that they be presented.
Appropriate references: in writing scientific papers, previous work and background in-
formation are always referenced. The most appropriate sources for references are other
journal articles (in peer-reviewed journals), books, and scientific databases. Internet refer-
ences are not appropriate for scientific papers. Here, “internet references” refer to private
websites or non-peer-reviewed information sites such as Wikipedia or other random web-
sites, even course academic sites. Books accessed online (such as through Google Scholar)
or journal articles found online are not considered internet references. Avoid using reg-
ular search engines; at minimum, replace your traditional internet search with a search
of scholar.google.com. If another reference cannot be found, the internet can be used
(sparingly, if ever).
In general, introductory college labs are not well suited to the type of scientific writing that
we will be exploring in this course. That said, in courses that focus more on quantitative analysis
where the outcome of each experiment is much more student and sample dependent the
inquiry approach and nature of the course opens up the possibility of writing our experimental
reports in the journal article style. It should be mentioned, however, that the scholarly papers
that you will be writing will, clearly, not have the same scope or depth as articles that are found
in journals.
1.3 To whom are we writing?
Possibly as important as the content of the paper is knowing who will be reading the paper
the Audience. Scholarly papers are always written with specific audiences in mind. These are
often determined by the type of publication; for example, the Journal of the American Chemical
Society is going to attract a more specialized audience than Scientific American. Scientists must
gear their writing toward their particular audience in order to have the greatest impact. The
main potential audiences of a scientific work are as follows, in order from most knowledgeable
to least:
Experts: these are people with expert knowledge of the general field of chemistry on
the whole and advanced knowledge of the specific sub-discipline (such as biochemistry,
computational chemistry, etc.). Journal articles and communications are generally written
for this audience.
Scientific: while not experts in the specific sub-discipline, or even the discipline, of the
author, these readers are experts in another scientific field; such as biology, physics, or
mathematics. Most scientists try to keep current on the major breakthroughs in other
field and will therefore read select journal articles, usually in the cross-discipline journals
such as Science and Nature.
Student: this is you. Writing for a student audience is designed to instruct individuals
starting at a basic level. Textbooks and lab manuals are good examples of writing done
with a student audience in mind.
6 · Undergraduate’s Guide to Writing Chemistry Papers
General: the general audience includes all readers, regardless of background, who are
interested in the field. Often readers will have no training whatsoever in the discipline, but
are interested in the topic. News articles and popular science articles are good examples
of works written for a general audience.
Each specific audience requires the writer to be aware of the limitations and the expectations
of the readers. There is very little assumed knowledge when we write for a general audience
meaning that we would be careful to explain all pertinent concepts and procedures in sufficient
detail. Conversely, it would be grossly inappropriate to include that same level of detail when
writing for a scientific audience, much more so for an Expert audience. Similarly, the emphasis
on motivation is likely to change for different audiences. A general audience is unlikely to care
about a new experimental approach that would be very exciting to other scientists.
There are two key differences between the type of writing discussed above and the realities
of classroom writing that we need to discuss so that you can attempt to look beyond them:
Your audience (instructors and graders) are already familiar with the research goals and
methods (and possibly even the expected results) of your experiments. Instead of trying to
convince an unfamiliar audience of the validity of your research, you may feel that all you
need to do is convince a familiar audience that you have properly performed an experiment
that they designed. Our goal, however, is to help you develop an understanding of writing
as it actually happens in the field of chemistry. Thus, you must imagine your instructor
and graders not as figures in a class, but as representatives of a larger group of readers who
belong to that particular academic area. Consequently, the vast majority of the writing
we will do will be for the scientific and expert audiences. Finding the correct balance
of information to present for your desired audience is one of the more nuanced parts of
scientific writing.
Unlike in an actual research setting in which you design your own research for some specific
purpose, in a classroom you (mostly) perform the experiments given to you because
that is what you are expected to do. This leads to the problem of understanding and
communicating the purpose or motivation for the experiment, which is discussed in the
next section.
1.4 Moving from student to scientist mindset
1.4.1 Writing doesn’t just mean writing
While it can be easy to think of writing as the act of putting pen to paper (or fingers to keyboard
keys), the writing process is preceded by analysis and research effort. In the case of scientific
writing, it is necessary that we first analyze the data that are recorded from experiment before
anything else. After all, without thoroughly understanding what we’ve achieved experimentally,
how can be begin to think about communicating those results with others?
The process of going from fully-analyzed experimental results to written paper is the main
objective of this work. Our goal is to develop the skills, ubiquitous to all genres of science and
engineering, necessary to go from experiment to paper.
Writing 1: Introduction to Scientific Writing · 7
1.4.2 Important things to keep in mind
Students new to science, and science writing, almost always make the same mistakes. One of
the most glaring is that they use overly complex terms/words (e.g. efficacious vs. effective;
proximal vs. close). When students start reading the scientific literature, or even just science
textbooks, many find themselves challenged by the depth and difficulty of the concepts being
presented. It is very easy to mistake the difficulty of the science with complexity in the writing.
In reality, scientists strive to write their complicated concepts in the most concise, precise, and
simple language possible.
Another challenge is the need to change how we think about the purpose of experimentation
and inquiry. Novices in science tend to think of the purpose of labs as “to get the right answer”
or “to properly replicate someone’s work.” As a result, lab reports tend to be thought of as
the vehicle for “this is what I got” (presenting raw data or numerical results) or “I got the
right/wrong answer because ...”. Journal articles, on the other hand, are about “I did this
research and this is my best understanding of why I got what I got. These results do (do not)
make sense in light of ...”. The bottom line is that numbers and tables are not answers; rather,
writing strong scientific papers requires one to develop the ability to research and explain the
chemistry. Consequently, most students find that they have to change how they think about
sources and references. The biggest mistake made by novice scientists is deciding on the meaning
of their data even before they’ve done their research and analyzed their results. This practice
often leads to a very immature relationship with the science and the scientific literature – these
students will oftentimes write their entire paper before ever engaging outside sources (if ever).
To prepare a paper that makes a strong argument, it is necessary to do enough research in the
literature to develop a thorough understanding of the chemistry behind the experiment, and to
use that knowledge when analyzing the data and explaining the results.
1.4.3 Objectives versus motivation
Before we start down the path to writing a scholarly paper, we must first clarify the purposes
behind what we are doing. For some students, it can be very difficult to correctly differentiate
between the different levels of motivation and objectives in their labs. For each experiment
performed we will look at the purpose, objective(s), and motivation.
Educational purpose: the (educational) purpose of a given experiment is the set of goals,
and topics, that the instructor intends for students to learn from any given exercise.
Clearly, this is not the motivation behind the student’s performance of the experiment
1
.
Lab objectives: these are the practical items that must be accomplished in the lab (create
a standard curve, digest a sample, etc.). These are the objectives that students will list
in their laboratory notebooks and they serve as the focus of the practical performance of
the experiment.
Motivation: the over-arching goal of the science that is being studied. Rarely, if ever, will
this be provided in the lab manual. The motivation for any given experiment is a good
scientific, or humanistic, reason why the topic or substances being studied are of interest.
1
For the sake of argument we’ll also assume that the student’s motivation is not simply to get an ‘A.’ Rather,
that the student is interested in learning.
8 · Undergraduate’s Guide to Writing Chemistry Papers
This is the Motivation that you will need to provide in the Introduction sections of your
papers or grant proposals.
Worked example: determining molecular size from a monomolecular film
Let’s consider an experiment in which the size of a molecule is determined by preparing a
monomolecular film of stearic acid at a water interface and analyze it on the three levels men-
tioned above.
Educational purpose: the main educational motives behind this lab might have been for
the students to gain basic lab skill proficiency (pipettes, balances), learn about calibration
(uncertainty, error), practice basic statistics (mean, standard deviation), gain confidence
in the lab environment, and develop skills working problems with unit conversions and
dimensional analysis.
Lab objectives: the two major objectives in the lab were to calibrate a micropipette and to
find the mass of stearic acid needed to make the monomolecular film using the calibrated
pipette.
Motivation: very little was discussed about why determining the size of molecule is so
important to study. In fact, there are many possible correct answers, and it is up to you
to determine one or more.
1.4.4 Difference between Data, Results, and Argument
As we continue discussing scientific writing some terms will keep repeating themselves: Data,
Results, and Argument. In the context of the scientific process, Data are the facts that will
be collected in the experimental (lab) setting or will be computed based on those experimental
measurements. In other words, data are facts about the physical world.
Results, on the other hand, are the outcomes or consequences of the data. Most experiments
involve recording substantial amounts of data, but there are usually few results. Finally, the final
outcome of doing science is to prepare and disseminate an Argument based on your experimental
results. In this context, your Argument is the takeaway message based on your results and a
thorough understanding, and discussion, of the principles guiding the science.
The process of doing science begins with observations (data) and culminates with delivering,
whether in writing or in person, a strong argument. It is this argument that will take its place
in the overall scientific discourse. We will discuss the components of logical arguments, and
how to organize them, in later chapters.
Continuing our worked example: molecular size determination
Continuing with our example of the determination of molecular size, consider the following
breakdown:
Data: volume measurements, concentrations, and surface areas are just some of the data
recorded in this experiment. None of these would likely appear anywhere in a paper
written about the experiment.
Writing 1: Introduction to Scientific Writing · 9
Results: the computed surface area of the molecule (with uncertainty) and how it compares
to other similarly recorded values are the results of the experiment.
Argument: it is impossible to decide what the argument would be without knowing the
results. The Argument is not about telling someone something that they already know;
rather, the goal of scientific communication is to Argue the meaning, validity, conse-
quences, and future of some experimental outcome(s) that have never before been observed.
This is the main goal of writing in the sciences.
1.5 This guide
The remainder of this Undergraduate’s Guide to Writing Chemistry Papers is organized to
guide a novice through the process of crafting a strong, scholarly paper. While some of the
conventions presented in this work are specific to papers in the field of Chemistry, most of
the practices discussed are applicable across the board in scientific disciplines. Moreover, the
concepts surrounding crafting a strong argument are similar, if not identical, to those in the
humanities and social sciences.
Writing Chapter 2 focuses on a superficial, yet crucial, aspect of science writing: crafting
effective and appropriate exhibits based on experimental data. These tables, figures, equations,
formulas, and schema will be those that are brought in the paper in order to showcase the result
of the experiment. They need to have a scientific look or no one will take your work seriously.
The next section (Writing Chapters 3 and 4) deals with how to craft a strong argument, and
how to organize that argument into the beating heart of an excellent paper. Then, in Writing
Chapter 5, the nuances and details of proper formatting for technical writing are discussed.
An in-depth discussion of research practices, the scientific literature, and bibliographies can be
found in Writing Chapter 6.
The remainder of the guide covers the other sections of scholarly papers (Introductions and
Experimentals, Writing Chapters 7 and 8), research proposals, and research presentations.
1.6 End-of-chapter assignment
1. In your own words, briefly list the differences between journal articles (i.e., scholarly
papers) and lab reports.
2. For which audience do you think it is easiest to write? Most difficult to write? Explain,
including a good reason for each of your choices.
3. Scientific writing is all about presenting a strong, cogent argument.
(a) How can the choice of sources used (and cited) in preparing a paper affect (positively
or negatively) the strength of the argument made? Explain briefly.
(b) Why is it important to ascertain the Results and Motivations behind an experiment
before conceiving the argument that you intend to use when writing your paper?
Explain briefly.
(c) How does understanding your intended audience affect the strength of your argu-
ment? Explain briefly.
10 · Undergraduate’s Guide to Writing Chemistry Papers
4. Consider one of the labs that you recently completed:
(a) Provide an analysis of the Educational purpose, Lab objectives, and Motivation of the
experiment.
(b) What Data did you collect? What are the Results?
WRITING 2: PREPARING APPROPRIATE EXHIBITS
2.1 Exhibits the start of developing a good argument
All writing is about making an argument a dynamic form of rhetoric designed to answer
questions and pose new ones. In scientific papers, these arguments serve as a mode of dissem-
inating the results of a few different types of scientific studies: developing new techniques and
methodologies, studying a novel and interesting system, refuting/confirming previous results or
beliefs, extending and refining previous work, and more. As scientific papers, especially ones
in a strictly regimented field like chemistry, all of these different papers would share many
structural and conventional similarities. Where these papers would differ substantially is in the
argument being made.
Before we can begin to identify what makes a strong argument that will be the subject of
later chapters we must first (a) figure out what our results mean and (b) format them as Ex-
hibits that would be useful for making a persuasive argument. Exhibits are the materials/items
in your paper that you are offering for explanation, analysis, or interpretation. Exhibits are the
grounds on which you will base the argument that you intend to make. As such, they are the
first things that we prepare after finishing the experiment.
In scientific papers, the exhibits will either be brought as part of the background material
or they will be used to present the results of the experiment. In the sciences, given the often
substantial amounts of data, exhibits will often be presented as figures or tables. The remainder
of this chapter will be dedicated to learning how to clearly communicate the results of an
experiment in Tables and Figures a highly underestimated, yet essential, skill to master.
2.1.1 Choosing the appropriate exhibit: tables and figures
If you’ve ever worried about how to successfully convince someone of your point of view, you
know that it is important to consider both (a) what you will say and (b) how you will say it.
Choosing the points that you want to make, and the depth to which you expound on them,
is extremely important. Say too much and you will lose your audience; say too little and you
won’t be convincing.
It is rarely appropriate to include all of the data that you collect in your experiment. One
of the earliest challenges that a scientist faces is knowing what data to present this a topic
that we will return to in later chapters. For now, we will focus on the equally-relevant task of
how to present the data that we’ve decided is important.
It can be tricky to know how and when to use a table, a figure, or a graph. Remember, the
goal is not to present large data sets large data sets are included in Supporting Information
11
12 · Undergraduate’s Guide to Writing Chemistry Papers
sections; rather, our objective is to prepare an exhibit that will convey a specific and detailed
message to the reader to support your argument. In that way, exhibits become efficient vehicles
of information and persuasion.
In general, the type of exhibit you choose depends on the data set that you have to com-
municate and your expected goals. Tables are best used for data that are qualitative, data
sets that include only a few points, or when you want to show the actual, specific values of
your measurements. Tables are most effective when seeing the details of the values will help
to convince the reader. Alternately, tables are the appropriate choice when it is necessary to
present each value for future reference. Figures, specifically graphs, are used more as a way to
visualize the trend of your data or the relationship between different sets of data. In this way,
figures have the ability to communicate concepts in far more efficient way than a table or prose.
In addition to the type of exhibit, you will need to decide how many figures or tables you
intend to use. In general, a figure is used to communicate one or two general ideas or trends;
conversely, tables can sometimes be used to arrange, present, and contrast vast amounts of
data. The best criterion for determining if a table or figure is needed is your overall goals. If
the figure or table adds to your readers’ understanding then it is necessary.
2.2 General formatting guidelines for exhibits
2.2.1 Numbering
Figures, Tables, and Equations, are all numbered sequentially (three separate sequences) in a
scholarly paper. These numbers (i.e., Table 1, Figure 1, Equation 1) are then used to refer
to the figure in the text where you describe the data. Tables, Figures, and Equations do not
“stand alone” in scientific papers; rather, they are exhibits that support the prose, but do not
replace appropriate discussion. To aid the reader in moving between the written discussion and
the exhibits, reference statements direct the reader to the appropriate exhibit.
Both tables and figures will have a label that will include its number. The label should
precede the title (above tables and below figures). Many authors use the good practice of
bolding the label (e.g., Table 1). Good reference statements include the label of the exhibit:
‘Table 3 shows the volume of water at various temperatures’ or ‘The molecular structures of
the products are depicted in Figure 12.’ Never use directional words, such as ‘below’ or ‘above’,
to describe the location of Tables and Figures, as editors may move the actual exhibits.
Note: in books or other chapter-based documents, exhibits are numbered by the chapter
(1.1, 1.2, etc.). For your papers, this type of numbering is not appropriate.
2.2.2 Tables have titles, figures have captions
Every Table must have an appropriate title, located above the table, that gives enough informa-
tion about the table such that, in the context of your report, an outside reader can understand
what the table is showing. An example of a Table title might be: Table 1: Volume of 1.0 g of
solvents at various temperatures.’
Figures do not have titles above them. The standard title that Microsoft Excel generates
is not appropriate for our purposes in both format (above the figure, large, and centered) and
content. Unlike tables, figures are captioned on the bottom, and the title is simply the first
part of the caption located below the figure and should be descriptive and complete.
Writing 2: Preparing Appropriate Exhibits · 13
A good figure caption elaborates on the information contained within the figure; it tells
you everything that you need to know in order to interpret the figure and get the ‘take-away’
message. Generally, a caption should be a few sentences long and follow the title. The title
should describe to the reader what is being presented, while the caption should tell the reader
any pertinent information that would be needed to interpret the data being presented. In the
case of a figure with multiple lines and no legend, the caption should also serve as the legend
e.g., ‘The trend before reaction (solid line) and after the reaction (dashed line)...’.
Captions are single-spaced (even in a double-spaced document) and have a smaller font size
than the body text. In this way, it is clear what is a caption and what is text of the paper.
Tables rarely have captions, unless there is a special note. In that case, a special character is
placed in the table (e.g., an asterisk () or dagger ()) and the note is placed below.
2.2.3 Placement in the document
It is best to try to place your table or figure close to the text that refers to it. However, it
is more essential that the exhibit be formatted properly rather than forcing it to be located
in a specific place. Editors have a habit of moving figures and tables to wherever they deem
appropriate (though they do keep them in order). Fortunately, since all figures and tables are
referred to in the text by number, it is not critical that they be placed on the same page as the
referring text. For the most part, figures and tables usually start or end a page or column (for
papers in the two-column format). Make sure that your exhibits are an appropriate size.
Conversely, equations are always located immediately before or after the text that first
describes them, as they are part of the prose and discussion. If the equation is mentioned again
later in the paper it is referenced, but not repeated.
2.3 Preparing effective figures
Figures can be a powerful way to present information in a visual medium. There are a few
types of Figures that you may find useful to include in your scholarly paper:
Non-data figures - You may be using or creating custom apparatuses for certain exper-
iments. Instead of tediously describing the schematic of an apparatus, it may be easier to
include a figure of the completed apparatus.
Crystal structures - A key step in synthesizing a compound is to determine its three-
dimensional structure by slowly crystallizing a sample of the compound and irradiating
it with x-rays. The results of that analysis are presented as a crystal structure.
Graphs - Graphs are one of the most important methods of presenting data in a visual
manner. They allow the author to present substantial amounts of data in a compact way
and allow the reader to perceive a trend in the information.
2.3.1 Considerations for preparing appropriate graphs
Consider the following when preparing graphs:
Which variable goes on which axis: There is no difference between calling a graph “pH
vs. volume” or “volume vs. pH” in both cases, the volume of base added is on the
14 · Undergraduate’s Guide to Writing Chemistry Papers
horizontal (x) axis. In a graph, the independent (manipulated) variable is plotted on the
horizontal axis and the dependent (responding) variable is plotted on the vertical axis
(y-axis).
Graph type - choose the correct type of graph for the type of data you are presenting.
While Excel’s default is the bar graph, which can be useful for presenting data to be
contrasted but that is not related. For data that shows a correlation (whether linear or
non-linear), it is appropriate to use an xy scatter plot.
In chemistry, most of your graphs will be xy scatter plots. Once you’ve chosen the correct
graph type, most software allows the user to choose how the points are linked:
No lines: most graphs will not have connected points. Since you only measured a few
data points in the experiment, it is misleading to connect them by lines – this implies
knowledge/data that is simply not there. Of course, it may still be appropriate to
include a line-of-best-fit (trendline) through the data.
Smoothed lines: example of exceptions to the previous rule are chromatograms, spec-
tra, and titration curves. Because these figures are made by taking regular and fre-
quent data points, they are plotted using a smooth line (without markers / points).
Axes - All axes need a label and units, unless the graphed quantity is unitless (such as
pH or absorbance). There is no need to start the plot at the origin (point 0,0) unless that
point is relevant to understanding the data (e.g., if you took a spectrum between 400 and
700 nm, the x-axis range should be 400-700). Also, never add a point at the origin unless
it was actually measured in the experiment.
Dividing lines - the default graphs produced by Excel include horizontal and/or vertical
lines at the major tick marks. Do not include gridlines on your graphs.
Legend - Legends are only useful if you are plotting more than one set of data on the
same graph. Even then, identifying the individual plots can be done in the caption (if
there aren’t too many). Legends are always necessary for graphs that present a large
number of data sets (more than 3), and many will include a legend for graphs with two
distinct data sets (a trendline and a data set are only one set of data).
Using space properly - You can resize the individual components of the graph to
maximize the use of the figure within your final document.
Color and shapes - For plots with multiple data sets, it is important to differentiate
between the them. Even if you print in color, it is advisable use different symbols for the
markers corresponding to different data sets. That said, make sure that data markers are
clear and reasonable shapes.
Trendlines - Where appropriate, for showing trends in the data, add a trendline and
display its equation on the graph.
Writing 2: Preparing Appropriate Exhibits · 15
2.3.2 Examples of well-executed figures
!
Figure'1:'UV-Vis!spectrum!of!carboxyfluorescein!in!methanol!(5!×!10
-3
!M,!1.00!cm!
cuvette).!The!maximum!absorbance!wavelength,!λ
max
,!is!located!a t!45 5 .0 !nm.!
!
!
!
!
Figure'2:'Cryoscopic!freezing!curve!for!cyclohexane!(K
f
!=!20.2!kg!K/mol).!!The!line ar !
trendline!(with!equation)!is!show n !fo r!th e !fre ez in g!p o rtio n !o f!th e ! curve!(between!
60!and!160!s),!and!the!extrapolated!freezing!point!is!6.7!
o
C.!!
!
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
380 430 480 530 580 630 680 730
Absorbance
Wavelength;(nm)
y"="-0.0013x"+"6.6914
2
3
4
5
6
7
8
9
10
0 50 100 150 200 250 300 350
Temperature"(⁰C)
Time"(s)
16 · Undergraduate’s Guide to Writing Chemistry Papers
2.3.3 Captions are extremely important and useful
It can be tempting to think of figure captions as just a label, often relegating them to an
afterthought. In reality, however, these captions are extremely important and contain a robust
amount of information.
The goal of a caption is to make an exhibit (figure) stand alone. An effective caption will
tell the reader exactly what they are looking at, guide them on how to read/understand the
figure, and draw conclusions about the implications of the figure. Look at some of the figures,
specifically their captions, in your general chemistry textbook. Does the level of detail surprise
you?
2.4 Preparing effective tables
While tables and figures are both useful for showing trends (though figures are much better at
this), tables are the preferable exhibit to use when it is important that the actual values be
accessible to the reader. That said, just because you are providing data to the reader, it is not
license to dump all of your raw data in one place. It is important that your tables have a good
balance of being complete, but also as streamlined as possible.
2.4.1 What needs to go into a table
A good table should make the data easier to digest. Therefore, it is not necessarily a good
choice to make a table that will have a single column or row of data that much could easily
be presented in the prose without being too hard to grasp.
Conversely, students often start off by mistakenly presenting too much in their tables, per-
haps in an effort to make them seem more significant. Masses of weighing jars, solids plus jars,
initial temperatures, final temperatures, etc., are all unnecessary Raw Data (the different types
of data will be discussed in a later chapter). For now, it suffices to mention that raw data is
never presented in papers (or even in Supporting Information for that matter).
2.4.2 Considerations for preparing appropriate tables
Most of the table formatting comes down to aesthetics. Consider the following important
guidelines for preparing your tables.
Header row(s) - All tables have a header row at the top. These headers explain the
type of data that will be found in a given column. Units are always included in the header
rows, sometimes in the second row. Often, header rows are bolded or shaded so that they
are distinguished from the other rows.
Breaking up the rows - Do not use gridlines in tables; this is simply not an appropriate
style. Do, however, use horizontal lines to create separation between the header row(s)
and the data below. Many people find it aesthetically pleasing to also place horizontal
lines above and below the table so that it is easily recognizable.
Row spacing - All rows should be the same height, with the exception of header rows
that can be taller than the others.
Writing 2: Preparing Appropriate Exhibits · 17
Column spacing - It is desirable that there be uniform space between the data in each
column, which means that column widths need to be adjusted to achieve even spacing.
Units and precision - Units are not included in the cells with the data. Instead, units
are included in the header row. Also, make sure to choose units that will lead to data
that is easy to read in the body of the table. For example, if small mass measurements
are being displayed, report the values in milligrams, and report the units (in the header
row) as mg or 10
3
g.
2.4.3 Example of a well-executed table
Table&6.1!"!Cereal!and!Fe!standard!addition!solutions
Solution&#
Volume&Cereal&
Solution&(mL)
Volume&Standard&
Fe&Solution&(mL)
Total&Volume&
(mL)
Average&
Absorbance
1 10 0 100 0.0012
2 10 1 100 0.0025
3 10 2 100 0.0065
4 10 5 100 0.0125
5 10 10 100 0.0250
6 10 15 100 0.0375
7 10 20 100 0.0500
8 10 25 100 0.0630
2.5 Equations, Chemicals, and Values
2.5.1 Which equations are relevant?
Chemical equations and formulas should be included if they are directly relevant to what you
are doing. Relevant chemical equations include those that describe the chemistry you are doing.
Numerical equations should only be used if they are relatively un-common, and are involved
with interpreting your data. Usually the entire derivation of an equation is not necessary, unless
the derivation itself is part of the work being presented.
2.5.2 Formatting chemical equations and the names of chemicals
In the sciences it is important the values, equations, and species be represented properly. For
example, in Biology, Latin names (binomial nomenclature) of species are always italicized.
Similar rules exist in the physical sciences for equations and values. Consider the following
guidelines:
Subscripts and superscripts - Make sure to properly format all chemical formulae with
appropriate subscripts and superscripts. Writing C2O42 is not acceptable (it should be
written as C
2
O
4
2
).
States of matter in equations - Where relevant, the states of matter need to be
included in chemical equations. We write Cu
2+
(aq), not Cu
2+
.
18 · Undergraduate’s Guide to Writing Chemistry Papers
Symbols - Always use the appropriate symbol for something. Use µ (not u) for micro,
for a normal equation arrow, and
*
)
for equilibrium. Note: means resonance and
cannot be used in place of equilibrium arrows.
Numbering - While equations do not need a title or caption, they should have an equa-
tion number. For a good example of how to format an equation with a label, consult the
examples in your lab manual.
Capitalization - Chemicals are not proper names. This means that the names of chem-
icals are not capitalized (HCl is hydrochloric acid, not Hydrochloric Acid).
Italics - In equations, only variables and constants are italicized. Letters representing
atoms should never be italicized. Carbon dioxide is written as CO
2
(g), not CO
2
.
2.5.3 Values and units
Given their tremendous importance in scientific communication, it is very important that values
be presented in an appropriate way. Consider the following:
Units - All measurements and values include a number and a unit. Always include a
space between the number and the unit. We write 10.5 m, not 10.5m.
Leading zero - Numbers less than 1.0 must be reported with a zero preceding the decimal
point (i.e., as 0.123, not .123). The decimal point and the digits that follow it are called
a mantissa, not a number.
Precision - Values should be presented in a clear way that shows how precisely the
number is known. For a value with a known uncertainty, this means only using the
appropriate significant figures in both the value and the uncertainty. The same applies to
statistical values.
Scientific notation - Numbers written in scientific notation are always written with ×10
(not the “E” or “e” that is the remnant of programming languages like Fortran).
Note: detailed instructions about typesetting equations and values, including how to streamline
the process, are presented in lab #1.
2.6 End-of-chapter assignment
1. Look at Writing Examples 2.1-2.5 in the pages that follow and explain the differences
between the Correct and Incorrect versions of the Tables, Equations, and Graphs. What
is the main point of each exhibit (i.e., what role does it play)?
2. Figure captions are necessary for making figures ‘stand alone.’ Explain how lack of detail
in captions negatively impacts the argument made in the paper.
3. We won’t be writing the traditional ‘lab reports’ (see Writing Chapter 1 for differences
between lab reports and scholarly papers). How might the amount and type of data
included in a table be different in a lab report versus a scholarly paper?
Writing 2: Preparing Appropriate Exhibits · 19
4. Consider a recent lab that required the preparation of at least one table. Prepare an
appropriate table using the information in this chapter. Don’t forget to include a title for
your table.
5. Consider a recent lab (or multiple labs) and prepare two appropriate figures: one spec-
trum/continuous and one calibration curve/trendline graph. Make sure to include de-
tailed, well-written captions for your figures, so that the reader will understand the main
take-away point of the exhibit without further explanation.
Writing Example 2.1: Table
Incorrect:
Table 3
Chocolate Brands
Type of
Chocolate
Peak Absorbance
Concentration
Dove
Milk
0.956
1.912 ppm
Hersheys
Milk
0.855
1.71 ppm
Lindt
Dark
0.869
1.738 ppm
Cadbury
Dark
0.981
1.962 ppm
Correct:
Table 4 - Quantifying levels of caffeine in chocolate.
Brand&
Type&
Peak&Absorbance&
Concentration&
(ppm)&
Dove&
Milk%
0.956%
1.912%
Hersheys&
Milk%
0.855%
1.710%
Lindt&
Dark%
0.869%
1.738%
Cadbury&
Dark%
0.981%
1.962%
!
Which one of these is more clear? Can you spot all of the difference between
the two versions? Note that this is just one way to properly format a table.
Writing Example 2.2: Equation
Incorrect:
P b2 + +2Cl −− > P bCl2
Correct:
Pb
2+
(aq) + 2Cl
(aq) PbCl
2
(s) (5.1)
Which one of these equations is better? Can you spot all of the changes?
20 · Undergraduate’s Guide to Writing Chemistry Papers
Writing Example 2.3: Reporting values with appropriate figures
Incorrect:
ε = 18713.45cm
1
M
1
± 518.3cm
1
M
1
Correct:
ε = (1.87
1
± 0.05
1
) × 10
4
cm
1
M
1
Which one of these equations is better? Can you spot all of the changes?
Writing Example 2.4: Graph (1) - Chromatogram
Incorrect:
-50
0
50
100
150
200
250
300
350
400
450
500
0 20 40 60
Series1
Series2
Correct:
0"
50"
100"
150"
200"
250"
300"
350"
400"
450"
10" 15" 20" 25" 30" 35" 40" 45"
Absorbance*(650*nm)*
Time*(minutes)*
Figure 3 - HPLC Chromatograms of chocolate content remaining in socks
after being washed with Washing Machine 1 (solid line) and Washing Ma-
chine 2 (dashed line). All absorbances were recorded at 650 nm. The
retention times for both machines are between 20 and 22 minutes.
Which one of these graphs is better? Can you spot all of the changes from
the generic Excel graph versus the properly formatted graph?
Writing 2: Preparing Appropriate Exhibits · 21
Writing Example 2.5: Graph (2) - Trend Line
Incorrect:
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0 20 40 60 80 100 120
Series1
Linear (Series1)
Correct:
y = 0.0032x - 0.0063
R² = 0.9919
0.2
0.22
0.24
0.26
0.28
0.3
0.32
0.34
0.36
60 70 80 90 100 110 120
Absorbance (650 nm)
Temperature (°C)
Figure 7 - Temperature dependance on washing machine efficiency. Peak
absorbance at the maximum wavelength, 650 nm, is plotted against the
temperature of washing.
Which one of these graphs is better? Can you spot all of the changes from
the generic Excel graph versus the properly formatted graph?
22 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 3: MAKING CLAIMS AND BUILDING AN ARGUMENT
3.1 Starting to think about argument
All writing is about making an argument. Each field has its conventions regarding the structural
and organizational details of research papers. Especially in a strictly regimented field like
chemistry, these conventions dictate how papers are arranged and composed. That said, the
common goal of all papers is that they be persuasive and, thereby, contribute meaningfully to
how people think about a field. This is the Argument, and it is the most important aspect of
writing a good paper.
Irrespective of the type of argument that you are going to make, your arguments will be
about making a strong claim that is based on grounds (evidence, data). The two most com-
mon types of grounds that we will rely on are: the exhibits that you prepare (based on your
experiments) and the research that you do (in journals and books). As a result, before you
can even begin to think about what you intend to argue in your paper, it is necessary to (a)
fully work-up your data, (b) prepare effective and convincing exhibits, and (c) research in the
literature to find information relating to your experiment and results. Additionally, it is highly
advisable that you take some time, after analyzing your data, before you start thinking about
writing your paper. Take a ‘step back’ and look at your results with fresh eyes this will help
you see global trends, impacts, and ramifications.
3.2 Planning your paper’s argument
As we’ve already said, all writing is about making an argument. In this context, “argument” can
be used interchangeably with the paper author’s “interpretation” of the results to an audience.
These claims are made based on the exhibits that we prepare, the results of the experiment,
and a thorough understanding of the background material (research in the literature).
An argument is made up of at least three key components: a claim, data (grounds), and
warrants. According to Toulmin, these three components make up the minimum logical argu-
ment that can be made. In addition, many arguments also contain additional, but not always
necessary, components: backing (for warrants), qualifiers, and counter-claims with rebuttals.
For now, we will focus on the first three components.
3.2.1 Crucial components of a logical argument
A Claim is an assertion being made about the nature of the world and is the cornerstone of
the argument. The entire purpose of writing your paper, and making your argument, is that
23
24 · Undergraduate’s Guide to Writing Chemistry Papers
the reader will agree with your claims. The claim is a conclusion, a statement of reasoned fact,
that is the focus of the paper. Your claim should answer a question such as what do I want to
prove? or what position do I want my audience to take? Note: this is not the same thing as
‘what value did you get?’ Rather, you should think more globally about what your data means
and what conclusions you can make based on them.
Think about what point(s) you want to make, not what paragraphs you want to write.
What is the one idea that you want a reader to walk away from your paper with?
Make sure that you’ve clarified exactly what you want to say. Don’t be wishy-washy and
make points that weaken others, that contradict other ideas that you’ve brought, or that
are not effective.
Avoid negative proofs: ‘A is bad, so B is better’ is not a strong claim.
Data (or grounds or evidence) are the foundation of the argument; these are the facts and
evidence that are brought to support the claim that is being made. In scientific inquiries, these
grounds can be from experimental data, conclusions made from an experiment, or cited research
in the literature. Your grounds are the answer to a question like what do I have to go on?”,
what data do I have to support my conclusions?”, or where must the audience start in order
to eventually agree with my claim? Think about the exhibits that you intend to display to the
reader:
How do you intend to present the data: in paragraph only, with the help of a table or
with a figure?
Exhibits (data, figures, and tables) are there for you to support your argument. Never
present extra data without a good reason. Which data need to be presented at all, and
what belongs in the Supporting Information?
There is usually a large disconnect between the claim that you intend to make and the
data/grounds that you bring to support it. Warrants are the links that are used to make
the connection between the claim and the data. Warrants, whether stated or implied, are
statements that are brought to give support to the data and show how they support the claim.
Warrants answer the question how do I get from evidence to claim?”.
We never assume that the reader will immediately jump to the same conclusions that we
do. For example, consider the claim that “candy X is bad” (claim) based on the fact that
it contains “20 g of sugar” (data). While this may seem like common sense (it is), in the
sciences it is important to bring warrants to bridge this gap.
More details about the types of warrants that are commonly used in writing can be found
later in this chapter (section 3.4).
3.2.2 Putting these components together
Together, a claim, data, and warrant comprise the smallest unit of an argument. Let’s consider
the following short argument that uses only these three elements:
Writing 3: Making Claims and Building an Argument · 25
Example: Sample argument with three components
“From the substantial amount of black smoke coming out of the 4th floor of
SCI (data), we conclude that the Metcalf Science Center is on fire (claim).
Smoke of this color is one of the most ubiquitous signs of fire
1
(warrant), so
we conclude that there is a fire in Metcalf (restated claim).”
Graphically, the argument in the above example can be effectively represented by the Graph-
ical Representation of Argument (GRA) diagram depicted in Figure W3.1. These diagrams are
an easy way to conceive, plan, and organize your argument before you start to write. By prepar-
ing a GRA before you start to write, you can make sure that all of the important components
are present and that your argument is cogent. Then, once you’ve decided that your argument
is sound, you can write your paper with confidence. This process will save you a substantial
amount of time in revisions and rewriting.
CH112%/%BUCWP% % Spring%2016%
%
3%
%
%
Data:%
Smoke%is%coming%out%of%SCI.%
Claim:%
SCI%is%on%fire.%
Warrant)statement:%
Since%smok e%is%on e%of%the%most%
ubiquitous%signs%of%fire,%
Figure)2:3GRA%Diagram%for%sample%argument%about%buildings%on%fire)
Figure 3.1: Graphical Representation of Argument (GRA) diagram for the example argu-
ment outlined about a fire in Metcalf. Note: this shape of GRA is not universal yours will
likely look very different.
3.2.3 Assembling graphical representations of arguments (GRAs)
Your graphical representation of argument (GRA) doesn’t need to look polished or fancy. The
most crucial thing is that your GRA contain the very specific components (claim, grounds,
warrants, etc.) that you want to use in your argument. Start by jotting down all of the
grounds (data, figures, tables) you might think about including in your paper. Consider using
a whiteboard or PostIt notes to make things easy to rearrange as you work you won’t get it
right on the first try.
Once you’ve surveyed all of your data, decide on the claim(s) that you intend to make. Jot
these down as well and add them to the board. Connect each claim that you want to make to
the data that is the foundation of that claim. It is conceivable that a single piece of data could
support multiple claims.
Next, you will need to prepare warrants to walk your audience from grounds to your claim.
There is no set number of warrants that are necessary for a given claim. Some claims are easy
to make and will require few warrant statements, while others may require many warrants. The
number is not important; rather, focus on making a strong argument that is convincing.
26 · Undergraduate’s Guide to Writing Chemistry Papers
Finally, it is important to note that the GRA diagram in Figure W3.1 was prepared before
the argument paragraph (in the example above) was written. Notice that once the argument is
diagramed, all that is left is to add filler words and work on the sentencing.
3.2.4 Identifying sources that strengthen your argument
One of the biggest mistakes made by novice writers is to decide on the argument that they intend
to make before they’ve done their research and analyzed their results. This practice often leads
to a very immature relationship with scientific literature — these students will oftentimes write
their entire paper before ever engaging outside sources (if ever). Those papers will rarely, if
ever, by very good.
A common question that instructors get is “how many sources do I need to cite?” Usually,
this question is meant to illicit a number of sources that will not result in a penalty for not
providing “at least 3 sources” (or however many the teacher is looking for). The problem is
that this approach is very counterproductive. Instead of the students engaging with the sources
(reading them and using the information as the basis of their papers), the students are just
looking to put citations at the end of the paper. The real answer to the aforementioned question,
as unsettling and frustrating as it may be, is that you should cite all of the sources that you
actually consulted when preparing your argument/paper, and no more. The most important
thing to remember is that sources cannot be relegated to an afterthought. These ‘sources’
are the heart and soul of the claim that you want to make. Details about the formatting of
references are found in Writing Chapter 6.
3.3 Additional components of strong arguments
Up to this point we’ve focused on the core components of logical arguments: claim, grounds, and
warrants. In his proposed structure for logical arguments, Toulmin discusses three additional
components that can be used to strengthen an argument: qualifiers, backing, and counter-
arguments and rebuttals.
When a warrant is sufficiently secure by itself, Backing is not necessary. We provide
backing for warrants to establish the reliability and relevance of the warrant statement, or
as a secondary source for the warrant. Use of backing is especially important when the leap
from data to claim is not small and would require several “steps.” In these cases, use backing
to establish a step-by-step chain that allows your audience to follow your argument without
developing questions about your logic.
To decide if your warrant needs backing, ask yourself the question is the move from
grounds to claim safe right now? or would everyone make the leap from evidence to
claim that I just made?”. If you answered either of these questions in the negative,
provide backing for your warrant(s) by citing other research papers, common knowledge,
or further experimental evidence.
Results from other students, fundamental concepts from the field of science, and method-
ologies are conceivable sources of backing.
Qualifiers are used in the wording of the claim in order to modify and limit the scope of
the claim. Words such as ‘usually’, ‘often’, and ‘under these circumstances’ can be used to limit
Writing 3: Making Claims and Building an Argument · 27
the claim. While this may seem to weaken the strength of the argument, it accomplishes the
opposite. By appropriately qualifying a claim, the author effectively identifies and states the
limits of their claim; this allows the reader to be confident about the reaches and limits of the
claims being made.
You should ask yourself what limitations do you see to the conclusion(s) that you are
trying to make? Remember: limitations are not bad, they are just the limits of the
claims.
Does your claim apply to all situations and samples? If not, how could you make this
clear in the most simple way. Phrases like of the ones we sampled or of this type can
qualify a claim.
Finally, Counter-claims and Rebuttals are used as a way to explore, and often explain,
the limits to the claim. A counter-claim is a possible objection to your claim that someone
reading your paper may have, and a rebuttal is your response to that counter-claim. Used
properly, counter-claims and their rebuttals are a way of acknowledging the types of objections
that someone reading the paper might bring up and then dismissing them. They are especially
useful for acknowledging exceptions that might otherwise invalidate the claim.
Ask yourself What objection might someone reading the paper have to my claims?”, “what
possibilities might interfere with my argument?”, or are there any exceptions that would
invalidate my claim?”.
Structure and back-up rebuttals to these possible counter claims as a way to reduce the
reader’s uncertainty in your claim.
Proper use of rebuttals will strengthen the initial claim as they provide a way of dispelling
possible detractors and acknowledging fundamental limitations of the claim. Often, delving
into a rebuttal will lead to another claim that will be argued and discussed using these same
components (evidence, warrants, etc.).
Let’s revisit the earlier, somewhat cheesy, example argument about a fire in Metcalf, and
add qualifiers, backing, and rebuttals:
Example: Sample argument with multiple components
“From the substantial amount of black smoke coming out of the 4th floor of
SCI (data), we conclude that it is likely (qualifier) that the Metcalf Science
Center is on fire (claim). Smoke of this color is one of the most ubiqui-
tous signs of fire
1
(warrant), which generally produces the smoke as fuel is
consumed
1,2
(backing). This implies that it is very likely (qualifier) that
there is a fire in Metcalf (restated claim), unless the smoke is being gener-
ated by a reaction in a research lab (counter-claim). Given, however, that
the building is currently empty for intersession (backing to rebuttal), it is
unlikely that the source of the fire is from a research group (rebuttal).”
28 · Undergraduate’s Guide to Writing Chemistry Papers
CH112%/%BUCWP% % Spring%2016%
%
2%
%
____%“Servers%making%a%dependable%wage%are%less%likely%to%seek%other%employ.”%
%
____%“Adding%a%service%charge%will%result%in%a%more-steady%income .”%
%
____%“Service%charges%eliminate%patrons%won dering%how%much %to%tip,%and%leads%to%more%equity%in%
service%compensation.”%
%
____%“Service%may%suffer%as%a%result.”%
%
b) Draw%a%GRA%Diagram%for%this%argument.%
%
Activity'#2:%%You%prepared%abridged%reports%before%arriving%at%class%today.%%Swap%abridged%reports%with%a%
student%sitting%next%to%you.%%Read%their%report%and%answer%the%following%questions.%
%
a) Was%the%author’s%Claim(s)%clear%to%you?%Make%a%note%of%the%Claim(s)%made.%
%
b) What%Data%did%the%author%bring%to%support%the%Claims?%%
%
c) Draw%a%GRA%Diagram%based%on%their%report.%%Did%they%include%any%unnecessary%components?%Were%
importan t%c o mponents%miss ing ? %
%
d) Present%your%constructive%and%supportive%feedback%to%you r%pee r.%
%
%
Data:%
Smoke%is%coming%out%of%SCI.%
(Qualified)'Claim:%
(it%is%likely%that)%SCI%is%on%fire.%
Warrant'statement:%
Since%smok e%is%on e%of%the% m o st%
ubiquitous%signs%of%fire,%
Backing'for'Warrant:%
Because%fires%generally%produce%smoke%
as%fuel% is%consumed,%
Counter-claim:%
unles s%the%smoke%is%coming%from%a%
research%lab%…%
Rebuttal:%
It%is%unlikely%that%the%smoke%is%coming%
from%a%research%group%
Backing:%
It%is%currently%intersession%and%the%
building%is%empty%
Figure'1:"GRA%Diagram%for%sample%argument%about%buildings%on%fire.'
Figure 3.2: Graphical Representation of Argument (GRA) diagram for the extended ex-
ample argument about a fire in Metcalf that contains all six elements of logical argument.
Note: this shape of GRA is not universal yours will likely look very different.
3.4 Common warrants used in scholarly papers
Warrants can take many forms and depend on the type of argument that you are trying to make.
In each case, these Warrants must be based in sound reason and be thoroughly researched. The
following is a list of some of the different types of warrants that you might consider using. While
this list is far from exhaustive, it is a good start for identifying the types of arguments that you
can make.
Argument based on generalization: this common form of reasoning involves making
the assumption that what is true for a well-chosen sample is likely to hold true for the
larger population or group. Alternately, this argument is that some things consistent with
the sample can be assumed about the larger group.
Backing for these types of warrants are discussions of things like the nature of the sampling
method used and why it is ok to generalize from the sample used to the larger group.
Argument based on sign or clue: this type of argument uses the notion that certain
types of evidence or trends (“clues”) are symptomatic of some bigger principle or situation.
Based on these well-established clues we can argue for some underlying cause or situation
that is not directly observed.
In making this type of argument it is necessary to include support (backing) for how the
clue is indicative of the underlying situation that you are implying is present.
Writing 3: Making Claims and Building an Argument · 29
Causal argument: in a causal argument the author argues that the results observed
are the result of, or affected by, a specific factor. While these arguments tend to be very
common in scientific studies, they also tend to be among the most complex. There are
two main dangers when using a causal argument: (1) conflating correlation with causation
and (2) mistakenly basing the argument in post hoc, ergo propter hoc (since it was this
way after the fact, it must be as a result of the fact).
Argument from authority: involves stating facts that are agreed upon by author-
itative sources and people. Relying on state-of-the-art methods, instrumentation, and
frameworks constitute argumentation from authority. In these cases, make sure to cite
the seminal work(s) in the field as backing for your warrant statements.
Argument from principle: similar to arguments from authority, these arguments are
based in the principles underlying the field of study and build upon prior work in the
field, even if there might not be complete agreement in the area.
Determining the warrants that you will use, and their accompanying Backing, will likely
require you to do research to locate, read, and digest papers and books in the field of study.
Citing papers, books, and review articles that you’ve used in cementing your argument is a
cornerstone of the scientific process.
3.5 Grounds choosing relevant data to include
It is rarely necessary to include all of the data that you collect in the lab in your Results section.
In fact, most times the actual data that is collected falls into the category of Raw Data. Raw
Data (also referred to as primary data) refers to any measurements that are collected in the lab
that have not been subjected to processing or manipulation. For example: masses of weighing
jars and solids, volumes on a burette, and initial and final temperatures, all fall in the category
of raw data.
Once the data have been processed you will have to further distinguish between the data that
are necessary to present in the body of the paper (as grounds or backing), and those which are
not necessary. Extensive, but relevant, data should be included in the Supporting Information.
To illustrate, consider the product of a synthesis: the only piece of data that will likely be
included in the argument is the percent yield. The raw masses are not relevant and should
not be given. Additionally, if you performed ten replicate mass measurements then only report
the mean (with associated confidence interval and N ); it is unnecessary, and inappropriate, to
report all ten measurements, the standard deviation, and the rest of the raw data in your paper.
3.6 Common mistakes
The most common mistake that novices make when beginning to write a paper is that they
don’t take the time to properly conceive the paper before they start “writing.” Rather, most
likely in the hopes of finishing quickly, they opt to start writing without properly organizing
their thoughts or outlining the major points that they intend to convey. As a result, the paper
can often be hard to read, lack flow, and omit critical components (which is not a good thing,
30 · Undergraduate’s Guide to Writing Chemistry Papers
especially if a grade is involved!).
Other common mistakes include:
Using data or statistics as warrant statements they definitely do not speak for them-
selves. Focus on the chemistry and concepts as warrants and backing (not statistics or
values). That said, under the right circumstances data can be used as backing to warrants.
Always include references in your GRA. Keep the sources that you use organized and
connected with the points that you are making this will make your paper-writing much
easier. In general, all backing that you did not generate in your experiment will need a
source.
Don’t start working on a GRA as if it will be your final product. It is best to allow
for flexibility in your workflow. Consider working with Post-it notes or on a whiteboard.
Once you’re happy with your GRA’s structure, prepare a final version. Microsoft Word,
or websites like Lucidchart, can be used to make a polised final product. Students with a
.edu email address can get a Free account at Lucidchart.com
3.7 End-of-chapter assignment
1. You intend to make a presentation to a group of restaurateurs to persuade them to change
their service/wait-staff policies to eliminate tipping and move to a fixed service charge.
Work in small groups (2-4 students) to answer the following questions.
(a) Classify each of the following statements as claims (C), data (D), warrant (W),
backing (B), etc.
“Servers may have an adverse reaction to the new policy.”
“Hiring new servers requires a large investment of time and money in things like
paperwork, training, and searching.”
“An 18% service charge should be added to patrons checks in lieu of tipping”
“High turnover of wait-staff is very costly.”
“Servers making a dependable wage are less likely to seek other employ.”
“Adding a service charge will result in a more-steady income.”
“Service charges eliminate patrons wondering how much to tip, and leads to
more equity in service compensation.”
“Service may suffer as a result.”
(b) Draw a GRA for this argument.
2. Consider a recent lab that was performed and go through the steps that you would take
in preparing a strong argument:
(a) Complete your data analysis. Tabulate the data and indicate where in the paper
(or Supporting Information) these would be included. List any exhibits (figures or
tables) that you prepared in your analysis. What exhibits, if any, would you use in
the actual paper? Explain.
Writing 3: Making Claims and Building an Argument · 31
(b) What is the objective of the experiment? Suggest a possible motivation.
(c) After completing your data analysis and preparing your exhibits, prepare a Graphical
Representation of Argument (GRA) for your paper using the questions from earlier
in this chapter to help your work. It is likely that you will need to do some research
in order to form your argument.
(d) What types of warrants or arguments did you employ? (section 3.4)
32 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 4: STRUCTURE OF ABRIDGED SCHOLARLY PAPERS
4.1 Structuring the argument into an abridged scholarly paper
There are a number of different types of arguments made by scientists and the conventions
that have been adopted for those modes of communication are directly related to the goal of
the argument. We’ll discuss two main types of arguments: (1) dissemination of results and (2)
solicitation of funds.
In order to make a persuasive argument about the importance and validity of their results,
scientists write peer-reviewed papers in a voice that gives the illusion that the scientist does
not affect the results, and they refrain (for the most part) from using the tenses that imply
subjectivity. Grant proposals, on the other hand, have a completely different purpose. They
are not meant to disseminate results; rather, they are for the purpose of having a granting
agency trust us (not just anyone) with money to do some (meaningful and useful) research.
As a result, grants are generally written in the present and future tenses, and use the active
voice. Often it can be the case that the even the best arguments can suffer because of lack of
understanding of the conventions of how to best make them.
For now, we’ll discuss the general framework for arguments being made when disseminating
results the journal article and we’ll return to writing grant proposals in a later chapter.
4.1.1 Framework of the typical journal article argument
The four main sections of a journal article (Introduction, Experimental/Methods, Results and
Discussion, and Conclusion) are not disjointed pieces (though they are written in such a way
that they can be read standalone). Properly done, a good journal article tells the story of a
scientific inquiry from inception to conclusion. Each section adds a new, different dimension to
the experiment and helps to reinforce the overall argument that the author hopes to make with
the paper.
Consider (a brief description of) the components that are found in each of the sections:
Introduction: the Introduction section of a journal article is where the author provides
the foundation for the experiment. (a) Introducing the research area: what is the topic?
Why is it important? Why should the reader care about the field/topic that you studied?
This is the Motivation that we discussed in Writing Chapter 1. (b) Identifying the gap:
what exactly did you study? Why? This is the Objective of the experiment. (c) Summary
of the literature/background pertinent to the study.
33
34 · Undergraduate’s Guide to Writing Chemistry Papers
Experimental: what you did in explicit detail enough that an expert reading the paper
should be able to reproduce the work, but without being overly verbose or including un-
necessary details. Demonstrating the validity and rigor of the methodology is a necessary
prerequisite to convincing the reader that the results have meaning. Note: the intended
audience will play a large role in determining the amount, and types, of details that are
necessary for an Experimental.
Results and Discussion: the argument what was found and what it means. This section
is, without a doubt, the most important section of the paper (which is why we will be
focusing on it first). This section includes all of the components of the GRA that you
outlined: claim, grounds, warrants, etc.
Conclusion: This section includes an extremely concise summary of what was accom-
plished in the experiment/study (usually only one sentence), in the context of the ob-
jectives. This is not a restatement of the claim. Additionally, the author makes claims
about the relative successes/failures of the work, suggests possible future work, and
most importantly explains the broader impacts that the results will have on the field
and society.
The remainder of this chapter will focus on the components of Abridged Scholarly Papers,
which do not include Introductions or Experimental sections. Abridged papers will present the
argument in its entirety, without the need to write the laborious sections that give the context
and background.
4.2 Results and Discussion Section
The Results and Discussion section is the cornerstone of the scholarly paper – this is where you
present your argument and claims, based on the results of your study and in the context of the
greater scientific objective. The Results section starts with restating the objective of the study,
and then data and exhibits are presented (the proper formatting of exhibit was discussed in a
previous chapter). Following this, you will explain your results (i.e., Discussion) to assert your
relative success or failure in relation to each of your stated goals. This is the climax of your
story it is crucial that you be clear, persuasive, and complete.
While it can sometimes be tempting to flood the reader with data, tables, and figures, the
most important thing to remember is that the Results and Discussion must be clear and readable.
The quality of the science is irrelevant if the reader cannot glean the relevant information. Use
GRAs and checklists (discussed later in this chapter) to organize your argument before beginning
to write your Results and Discussion.
4.2.1 Content and Organization
Use checklists and GRAs to organize your thoughts, and your points, before you start writing
your Results and Discussion. A detailed checklist, with the actual points that you intend to
make, will help you to quickly and effectively write your paper. The beginning of the Results and
Discussion section, or each subsection in the Results and Discussion, is a concise restatement
of the objective of the study (in no more than one or two sentences). The specific, take-away
results of the experiment and their context are then presented (claims).
Writing 4: Structure of Abridged Scholarly Papers · 35
To present your results you will need to be bring select exhibits (grounds; see previous
chapters for how to determine if data is relevant to be included and the best methods for
displaying it) and explain the reader what they should learn and infer from them. Captions of
figures are tremendously important for getting points across, but the prose in which exhibits
are discussed is paramount.
The most important part of the Discussion is to make a clear argument about the meaning
and scope of your Results (warrants). If the objectives of your study were partially met, or the
hypothesis was supported with exceptions, this should also be explained clearly and reasoned
well (including qualifiers, counter-claims, and rebuttals). Make sure that your explanations of
the chemistry concepts are clear, and supported by references to scientific literature, in order to
shape a strong argument about your Results. Remember, experiments are not about accuracy
and precision of results – these are details. Your paper is about the chemistry of the experiment.
Make sure that your argument makes sense and that the points (warrants) you make are done
in support of your argument.
Additionally, one thing to keep in mind is that most of the experiments that you will perform
in a four or eight hour lab session are of limited scope and length. Consequently, the types of
claims and the amount of results that you have will tend to be minimal.
Finally, remember that the Results and Discussion is not the section for overarching gener-
alizations about the overall success or failure of the experiment. Those types of statements and
discussions are left for the Conclusion section.
4.2.2 Format and Style
The Results and Discussion section will consist of paragraphs and relevant exhibits (tables,
figures, and equations). A Results and Discussion section that consists uniquely of tables and
figures is not acceptable.
While there is no page limit for this section, it is best to keep the Results and Discussion
as concise as possible. If your experiment contains multiple (more than two) central results,
then you may find it useful to break your Results and Discussion into smaller subsections. Each
subsection is about a specific claim, and should start with a heading (e.g., Spectral Data or
Stability Results). Do not use “Part ...” in your headings.
The Discussion consists almost entirely of prose that is written in the past tense and passive
voice. While the Discussion should be persuasive (anyone would come to the same conclusions
based on your data), it must also be objective. Relevant chemical equations, tables, and figures
are also included where they are needed to explain each Result. Sample calculations never
appear in the Results and Discussion (they should be shown in your Supporting Information
section, which appears after the References; never reference, or direct your reader to, your
Supporting Information in your paper).
4.2.3 Proper use of Exhibits
All exhibits tables, figures, and equations must be referenced by number in the text of
the paper. Do not use directional indicators, such as above or below, as the exact location
of exhibits can often change before publication. Additionally, make sure that all exhibits are
properly formatted, and that all figures have meaningful captions (see Writing Chapter 2 for
additional details).
36 · Undergraduate’s Guide to Writing Chemistry Papers
4.2.4 Voice and Tenses
Substantial portions of chemistry papers are written in the passive voice and the past tense this
gives the impression that the results of the experiment are not dependent on the person actually
performing the work, but that they would be the same if anyone would do the experiment. The
Results and Discussion section will contain both the past and present tenses, where appropriate.
The work done in the past should be described in the past tense (“the structure was determined
by ...”), while the claims that are gleaned from the research are in the present tense (“the
structure indicates that ...”).
4.2.5 Common Mistakes
Always start your Results and Discussion with a restatement of the objective of the study.
This is necessary because each section of a paper needs to stand alone.
Make sure to be concise and avoid being overly verbose. While it is important that the
Results and Discussion ‘flow’, it is important to keep it relatively short and to the point.
Start by making a GRA and expanding it into a checklist.
Students often present too much data: discussing too much in a single paragraph, overuse
of Tables and Figures, presenting too much raw data, presenting the same data in multiple
places (it is unnecessary to present every datum from a table in the text).
Writing Example 4.1: Raw data
Incorrect:
“The density of the products were recorded and found to be 1.2
g/mL, 1.1 g/mL, 1.2 g/mL, 1.0 g/mL, 0.9 g/mL, 1.3 g/mL, 1.4
g/mL, 1.2 g/mL, 1.3 g/mL, 1.2 g/mL. The average was found
to be 1.2
1
g/mL and the 95% confidence interval was 1.2
1
±0.1
1
g/mL.”
Correct:
“The average density of the product (N = 10) was determined
to be 1.2
1
± 0.1
1
g/mL.”
For all of the reasons mentioned previously, the second statement is far
superior to the first statement. In the Data Analysis section of your Ex-
perimental section you would report that all confidence intervals were com-
puted at the 95% confidence limit.
Do not summarize the lab in the Results and Discussion. A summary of the relative
success or failure, in brief, is made at the beginning of the Conclusion section. The
Discussion section should involve details rather than vague generalizations. That said, it
is important to make sure that you state whether or not your results support your original
hypothesis.
Make sure that your assertions are supported by properly-formatted references to the
scientific literature. Also, be careful to make assertions that make sense! For example:
Writing 4: Structure of Abridged Scholarly Papers · 37
losing product during filtration is not a good source of uncertainty when the experimental
value is larger than expected. (Remember: instructors and lab partners are not sources
of uncertainty/error; human error isn’t a real thing in scientific writing.)
Never reference your supporting information; if it is critical information then it needs to
be presented in your paper.
Avoid using an informal tone in a Results and Discussion. In an effort to be persuasive
you may be tempted to flirt with other moods do not. It is more important that the
Results and Discussion remains objective.
4.3 Conclusion Section
The conclusion ties together the entire paper. To this end, you should not simply duplicate
statements made in the Discussion, rather you should present a brief overview of the whole
study, with a summary of the main claims/findings, and conclusions of the study, particularly
the larger ramifications.
4.3.1 Content and Organization
There are three major components of the Conclusion section:
1. Summary: Briefly summarize the main conclusions of the lab, including a brief restatement
of the objective/goal of the study and claims. This is not a repetition of the first part of
the Results and Discussion; rather, this is a summary of what was achieved and what the
reader should take away. Were the study’s objectives achieved? If so, to what degree?
If not, how unsuccessful was the experiment? Note: there is no need to re-prove your
claims; you did that in the Results and Discussion.
2. Future Developments: Suggest, and where appropriate explain, any future development
or work that could be done as a result of this study. In other words, “what comes next?”
These ideas and suggestions need not be grandiose. Actually, better solutions are simple
ones that are easy to implement, but they should have a relatively large potential for
impact and be consistent with the results of the experiment. This is the place to think
about future research that could be done based on the results or the methodology.
3. Global ramifications: Overall, what was accomplished? “What did you learn?” Note:
what you learned does not refer to the educational purpose of the lab (titrations, UV/Vis,
AAS, etc.). This is referring to what was learned from the experiment (e.g., the best cereal
to eat for iron is ...). How does this change the world? In general, considering that these
are undergraduate labs, there is likely to not be any world-changing global ramifications.
That is ok. Discuss the ramifications without resorting to hyperbole.
Finally, it is very possible, especially if the experiment was not successful (or had limited
success), that there are no global ramifications to the work. In this case, focus more on
suggesting methods of improvement.
38 · Undergraduate’s Guide to Writing Chemistry Papers
4.3.2 Format and Style
Voice and Tense
Unlike the Results and Discussion section that was written in the present tene (claims) and past
tense and passive voice (results), the Conclusion discusses the present and the future. Keep the
following in mind:
Summary: The summary is about an experiment that was performed in the past. This
component should be written in the past tense and passive voice (like the rest of the
paper).
Future development: This component is referring to work that may, or may not, be done
in the future. Additionally, since it is yet to be determined whether or not someone
will actually make those improvements, it is not appropriate to use the Indicative Mood.
Rather, the author should write these statements either in the Conditional Mood or the
Subjunctive Mood. The Conditional mood is marked by the use of the words might, could,
and would. In either case, the goal is to make it clear that these are suggestions that may
never actually come to fruition. This portion is still written in the third person.
Global ramifications: These statements will either speak about how the study will impact
the future of the field or how they have changed the current state-of-the-art. These
statements should be written in the future indicative and present indicative, respectively.
In chemistry, most of scholarly papers are written in the third person and passive voice. This
is done to give the impression of objectivity that the results, and conclusions that are drawn
from them, would be achieved had anyone done the experiment. As a result, it is important to
remember to avoid writing your conclusions in the first person (I or we).
Length and Depth
The Conclusion is the shortest part of the research paper – a single paragraph (sometimes two),
that should not significantly exceed half of a page. Conclusions will (almost) never present new
exhibits.
4.3.3 Common Mistakes
Suggestions for improvement and future work should be possible and reasonable. Also,
since it is not clear that the experiment would ever be repeated, future development should
be discussed in a passive voice (suggestion), rather than a definite statement about the
future. Make sure to cite your sources.
Don’t restate your claims or re-argue your position you did that in the Results and
Discussion. The Conclusion is about summarizing the paper and looking to the future.
Avoid going overboard with the familiar and personal voice in the Conclusion; make sure
that the Conclusion still reads like part of a scientific paper and not a personal notebook.
Writing 4: Structure of Abridged Scholarly Papers · 39
Example Conclusion
Incorrect:
“Completely abandoning the field and reinventing the wheel
will lead to better precision.”
Correct:
“Ensuring location consistency in the collection at the Charles
River would likely decrease variation introduced by sampling.”
While the first statement may sometimes be true, it does not give any
real constructive suggestion (your assertions and suggestions should be
specific). Additionally, it is written in a manner that leads the reader to
believe that the author will definitely implement their future development
work.
4.4 Putting together the complete, abridged paper
4.4.1 Each section stands alone
Keep in mind that scientists rarely read a paper from start to finish. Instead, we usually are
reading a paper for a very specific purpose. For example: to extract some necessary data
(Results), to glean some understanding of an important/interesting phenomenon (Discussion),
or to learn how to perform a specific procedure/synthesis (Experimental). Notice that none of
those reasons require the reader to agonize over the entire paper; rather, we will often only read
the section(s) of the paper that are necessary for our purposes.
Consequently, it is important that each section be able to stand alone. The Introduction
section will always contain the motivation and objective of an experiment, as well as a de-
scription of the general approach taken to solve the scientific problem. The Conclusion section
begins by summarizing the overall success/failure of the experiment. All that remains is to pro-
vide a brief and concise summary of the problem to be solved at the beginning of the Results
and Discussion. The captions of tables and figures (or, more accurately, a rephrasing of them)
appear in the text as the table/figure is being discussed.
4.4.2 Appropriate format for research papers
Your papers will usually be comprised of two parts: the paper (including references) and sup-
porting information. The paper will always start on a new page with an appropriate title, the
name(s) of the author(s), and the author’s affiliation. References do not start a new page, but
follow immediately the conclusion of the paper.
Write your title last, once you’ve completely digested the message of your paper. The title
must be specific and informative; the goal of a title is to inform the potential reader about the
content of the paper and to entice them to read the paper. Additionally, your name, the name
of your instructor (teaching assistant), the name of your collaborators (i.e., lab partners), and
the name and address of your institution are included. An asterisk (*) next your name denotes
that you are the corresponding author.
The length of the title is often indicative of its quality. Titles that are too short (3-4 words)
almost always lack specificity, while overly long titles (15+ words) are equally inappropriate.
40 · Undergraduate’s Guide to Writing Chemistry Papers
Specificity is key in titles; make sure to use terms that are specific to your paper. It is best
to avoid using flowery terms and unnecessary filler words. Symbols and equations are almost
never permitted in titles. Additionally, it is proper to omit the “The” at the beginning of a
title.
In journals, authors use the Supporting Information (also referred to as the Supplemental
Material) to include extra material that they feel that the reader might be interested in reading,
but is tangental to the actual article. Typical items in the Supporting Information are (1) sample
calculations (rarely found in journal articles, but you must always give one per calculation in
your labs), (2) spectral data that is not a central result to the paper, (3) unpublished results
that could shed light on your work, and (4) anything that you feel will clarify the reader’s
understanding of the work. For our purposes, the supporting information sections of your
papers will contain the full set of data analysis from your experiment and the answers to any
Questions for Thought. The supporting information (data analysis) must be complete and in
the appropriate order. Exhibits (tables, figures) that will be presented in the paper also appear,
in the appropriate place, in the supporting information.
Example Title
Incorrect:
Acid-Base Titrations
Correct:
Determination of the pKa of Acetic Acid via Poten-
tiometric Titration with Strong Acid
Binyomin Abrams* and Sample Name
Department of Chemistry, Boston University, Boston MA
In the case of the second title, it is much more clear exactly what exper-
iment was performed and what the reader can expect to read. Note: the
asterisk (*) is used to indicate the corresponding author.
4.4.3 Overall considerations for abridged papers
There is a lot to consider when approaching your abridged paper. Consider the following:
Start with a good GRA (see Writing Chapter 3) for your paper. Use this GRA when
planning your checklist.
A good checklist for your paper is necessary in order to produce a high quality paper.
Make sure that your checklist is consistent with your GRA.
Focus on developing a concise paper that has a good flow and a strong argument.
Make sure to follow the guidelines about what to avoid in scientific writing.
After completeness, focus on the correct use of voice and tense. To achieve a paper with a
strong argument you will need to write in the proper tenses. Make sure that each section
stands alone.
Writing 4: Structure of Abridged Scholarly Papers · 41
Make sure that all exhibits are referred to by number in the text of the paper, and that
they follow proper formatting (see Writing Chapter 2).
Always remember to look over your work with a critical eye for the ‘common mistakes.’
They are so named because they are very common!
4.5 Using a checklist to effectively outline a strong argument
When writing a paper, it is natural to develop a mental checklist of the things that we want to
include. It has been our experience that explicitly writing down this checklist, before starting
to write the paper, is one of the most instrumental steps towards ensuring that the paper is
complete and well-organized. Moreover, using the checklist as a way to outline the argument
that you want to make is the best way to ensure that your paper is cogent, well-organized, and
able to properly convey the information that you want to present in an order that makes sense.
Start with a GRA and expand it into a checklist.
For the papers that we will write, that strongly mimic the journal article style, a checklist
that starts with the components included in the Supporting Information, and continues section-
by-section through the rest of the paper, is ideal.
4.5.1 Where does it belong?
Before preparing our checklist, we need to identify the components that could/should be in-
cluded in our paper. Determining the complete list of data and exhibits is trivial; conversely,
determining the appropriate place to include an item in our paper can often times be very
difficult. Consider the following:
First, make sure that your Supporting Information
2
contains one complete copy of your
data analysis. Your teaching assistant will be grading the technical merits of your analysis
from your Supporting Information, so it is important that your analysis be complete with
sample calculations, tables of relevant data, and the figures that you construct. While
the supporting information comes at the end of a paper, it is the first thing that you’ll
prepare (so put it first in your checklist).
Other than in the Supporting Information, sample calculations should not be found any-
where else in your papers. Additionally, raw data will not be included in your paper.
The Results and Discussion section will contain a subset of the exhibits that your pre-
pared during your analysis. The most important part of the Results and Discussion are
the claims that you will make, supported by these exhibits (i.e., grounds) and warrants
(discussion of the chemical principles that underlie the results and the implications of
these results).
Every exhibit that you discuss in the body of your paper needs to be included, but not all
exhibits that you prepare should be discussed. Make sure that your Results and Discussion
section only includes exhibits and data that are directly relevant to the claims you are
2
In the ‘real’ science writing world, the Supporting Information contains everything that you (or, more
importantly, your reader/instructor) find informative, but that does not have a home in any of the other sections.
42 · Undergraduate’s Guide to Writing Chemistry Papers
making. Ask yourself: “does this information get across my message to my reader? or,
does including this obfuscate the point that I need to make?”. For a novice writer, this
is possibly one of the most difficult parts of writing scientific papers. Often, students will
choose to include substantially more than is appropriate. Not only does this waste their
time, but it also detracts from the strength of the paper. Exhibits that are important for
sub-claims are also included in the body of the paper.
4.5.2 Preparing a checklist for your paper: a step-by-step guide
Especially when writing a paper in the sciences, there is a long process that precedes the writing
of a paper or even a checklist. Novices who “jump” straight into writing their paper will usually
end up wasting a tremendous amount of time and rewriting large sections, if not the entirety,
of their paper. Consider the following approach that experts use when writing their research
papers:
1. Data Analysis: Whether the subject of your paper is Dante or chemistry, you need to know
what you intend to write about before you start writing! Start by completely analyzing
your data and preparing any relevant exhibits (figures and tables). Only once you’ve
sorted out the results of your experiment can you begin to ponder what you intend to
argue.
2. Plan your argument and do your research: These two tasks, planning an argument and
doing research, are inextricable. Answer the questions outlined earlier in this chapter to
guide your research and construct a GRA. Make sure to keep track of the sources that
you used in your research. It is helpful to label the components based on their role (claim,
grounds, etc.).
3. Parse your data and prepare a checklist: Expand your GRA into a checklist of items that
you intend to include in your paper (in an order that makes sense).
4. Be mindful of section-specific details: Results and Discussion sections start with a restate-
ment of the objective(s) of the study, Conclusions start with a summary of the success of
the study, etc.
4.6 Qualities of a good checklist
Consider the sample checklist that is included at the end of the chapter. This checklist is one
that might have been used for writing a paper reporting on an experiment to measure the sugar
concentration of sodas by their densities. Notice several important features of the checklist at
the end of this chapter:
Be very specific! Don’t just write “motivation” or “reason for studying sugar”; rather,
actually list the specific items that you intend to write about.
It is very terse. This is not the paper, or even the beginnings of the paper. Rather, it
is a comprehensive checklist of all of the things that you would want to have in the final
draft.
Writing 4: Structure of Abridged Scholarly Papers · 43
While colloquial writing isn’t acceptable for writing scholarly papers (nor is the use of the
first person), it is perfectly acceptable for your checklist to be colloquial.
Include citations and bibliographical information to indicate the items that were the result
of your research and, hence, need citations to the chemical literature.
The Supporting Information is first! The first place to start when writing a paper is to
fully work up the data.
The importance of checklists cannot be overemphasized they are an indispensable tool for
streamlining workflow and organizing the scientific content of papers. Traditionally, the grades
on papers tend to be much lower when students do not produce high-quality checklists.
4.7 End-of-chapter assignment
1. Differences between lab reports and scholarly papers. Based on the information in this
chapter:
(a) List two ways that your Results and Discussions are different in scholarly papers
than in lab reports that you’ve written.
(b) List two ways that your Conclusion sections will be different in scholarly papers.
2. Revisit the example of the “fire in Metcalf from Writing Chapter 3. Prepare a checklist
for an abridged paper based on the GRA that was presented. You will need to add more
components than those that were listed.
3. Consider a recent lab that was performed and go through the steps that you would take
in preparing a strong argument and the checklist you would use to write a paper:
(a) Complete your data analysis. Tabulate the data and indicate where in the paper
(or Supporting Information) these would be included. List any exhibits (figures or
tables) that you prepared in your analysis. What exhibits, if any, would you use in
the actual paper? Explain.
(b) What is the objective of the experiment? Suggest a possible motivation.
(c) After completing your data analysis and preparing your exhibits, prepare a Graphical
Representation of Argument (GRA) for your paper using the questions from Writing
Chapter 3 to help your work. It is likely that you will need to do some research in
order to answer these questions.
(d) Prepare a checklist, of the style described above, that you could use to write a
strong paper based on your GRA. Remember that your goal is to prepare a strong,
and well-organized, argument.
44 · Undergraduate’s Guide to Writing Chemistry Papers
Example: Sample Checklist
Experiment #42 Partner: W. Eirdo Due date: 1/1/1900
Supporting Information
Table with masses, volumes, densities, and %w/w sugar of the standards.
Table containing the masses and volumes of the soda samples used; densities
of soda samples; and concentrations from standard curve
Figure: Standard curve for %w/w sucrose vs. density
Table for soda unknowns: average (±s) density, average (±s) %w/w sugar,
%error from known value
Sample calculations for (a) the densities from mass and volume and (b) the
statistics for multiple trials.
Results and Discussion
Goal to determine worst soda from health point of view (state objective)
Mountain Dew is 15% sugar - the largest concentration (data)
Which makes it the worst soda to drink (claim)
... of the sodas analyzed in this study (qualifier)
Exhibit #1 is figure of standard curve for %w/w sucrose vs. density (warrant).
It is used to show consistency in sampling and precision (backing).
Exhibit #2 is a table for soda unknowns: average (±s) density, average (±s)
%w/w sugar, %error from known value (grounds).
Fresca has the least - none; control study (warrant)
Complete χ
2
analysis shows consistency with manufacturer claims (rebuttal),
even though some beverages had larger deviations than others (counter-claim).
Method is reliable: linearity in the curve, R
2
of >> 0.9995 (*) (rebuttal)
Reason why method works best for some beverages and not others is the other
components in the soda contribute to the density and are not consistent (*)
(rebuttal, new claim)
Conclusion
Method successfully determines %sugar that matched manufacturer’s claims
(restate claim)
Broader impact: discontinue drinking Mountain Dew because it is 50% worse
than other sodas
Also, this method is effective for low-cost field evaluations of sugar solutions
(secondary claim, future directions)
WRITING 5: CONVENTIONS OF SCIENTIFIC WRITING
5.1 What to avoid in scientific writing
Having discussed the general type of writing that we will be doing, polished our use of exhibits,
and outlined our paper with a GRA and a checklist, we are now ready to begin writing our
scholarly papers (in the Journal Article style). Before you begin writing, first consider the
following general guidelines about what to avoid when writing scientific papers:
1. Avoid colloquialisms: it is very tempting to write the same way we speak. Colloquialisms
can be broken down into a few categories: inappropriate words, such as gonna, y’all, or ain’t);
idiomatic phrases, like “since the dawn of time” or “as blind as a bat;” aphorisms, iconic
sayings, like “There’s more than one way to skin a cat;” and profanity are all inappropriate in
scientific writing. Additionally, make sure to avoid personifications of chemicals, equipment,
and results; never use phrases like “tells us”, “prefers”, “says”, “wants to be”, or similar
anthropomorphizations. These turns of phrase, while appropriate in other forms of writing,
must not be used in scientific or technical writing.
2. Avoid “Lab Speak.” In general, lab speak refers to language that identifies the writing
as being for a course in college, rather than a body of scientific information. Examples of
inappropriate language include: in the lab, in this experiment, the student, and in the procedure.
In many cases, the offending clause can be omitted, or in some cases replaced with more
appropriate versions such as ‘in the present work.’
3. Avoid using inappropriate (or made-up) words and phrases. Students new to
science, and science writing, almost always make the same mistake: they use inappropriate
terms and phrases in their writing. Sometimes it is just that they are using a term that sounds
like should exist (e.g., “massing” is not thing). Other times they conflate terms or use different
things interchangeably; words like substantial and significant, related and correlated, all have
very well-defined meanings in the scientific world, and it is important that you use the correct
terms. In either case, it is important to choose the most appropriate words when writing in the
sciences words that are (nearly) synonymous in other forms of writing can often have very
different meanings in science.
4. Never use direct quotes. Don’t cite for the sake of citing and never use direct quotes. Your
citations are for information that you used from other sources and are brought to support your
assertions and your argument. That said, always cite the sources that you used in researching
45
46 · Undergraduate’s Guide to Writing Chemistry Papers
and papering your paper.
5. Be specific, never vague, and avoid using overly grandiose statements. It is important
to be very specific in scientific writing. Words such as numerous, incredibly, enumerable, and
essential, are usually inappropriate.
Writing Example 5.1: Overly General Statements
Incorrect:
“Potassium Aluminum Sulfate is an incredibly useful inorganic
white crystalline compound possessing numerous purposes both
in regards to ecological sustainability and also everyday usage.”
Correct:
“Potassium aluminum sulfate has uses in water purification
1
,
paper sizing
2
, leather tanning
3
, mordant dyeing
4
, synthesizing
baking powder and antiperspirants
5
, and usage as a medical
astringent.
6
Notice that the first version is extremely vague. The second version enu-
merates the uses of Alum, instead of making unfounded, grandiose claims.
Also, each claim in the second version is supported by an appropriate
reference to the literature.
6. Refrain from using anecdotal statements. It is not uncommon for students who are
new to scientific writing to make statements that begin with “it is well known that” or “most
people know.” When writing for a scientific audience, however, these types of statements are
inappropriate. Consider the following: if they are so well known, then you need not say them
at all. If, however, they are important enough to say, then you should make sure to support
your claim/statement with appropriate references. In general, statements of previously-known
scientific fact that are necessary to support your warrants should be supported by references to
authoritative sources.
Writing Example 5.2: Anecdotal and general statements
Incorrect:
“Despite this lack of confidence, the consumption of vitamin
supplements is very widespread.”
Correct:
“Recent data
1
indicate that 90% of the population consumes a
vitamin supplement regularly.”
The first statement is extremely vague and is not supported by any sources.
The corrected version gives the same information (that most people con-
sume vitamin supplements), but is also very specific and is supported by
a reference.
7. Run-on sentences are somewhat difficult to avoid while we are striving to be concise in our
writing. That said, a quick re-read through your paper should help you to identify sentences
Writing 5: Conventions of Scientific Writing · 47
that are too long and complex. Similarly, if you have two or three short sentences about a single
point, try to combine them into a single, well-written, sentence.
8. Avoid being repetitive or indirect in your writing. Say what you intend in a direct
and clear manner. Use GRAs and checklists to organize your writing before you start, and make
sure that the points you want to make are complete. Avoid repeating the same points or details,
except where it is stylistically required such as summarizing the outcome at the beginning
of a Conclusion or restating the objective at the beginning of the Results and Discussion.
Finally, use references to the literature in order to give readers a place to go for more detail
and background.
9. Avoid overly complicated words and phrases. When new students start reading
scientific papers, or even just science textbooks, they find themselves challenged by the depth
and difficulty of the concepts being presented. It is very easy to mistake the difficulty of the
science with complexity in the writing. In reality, scientists strive to write their complicated
concepts in the most concise, precise, and simple language possible.
5.2 Best practices in scientific writing
Knowing what to avoid is only half of the battle – it is more important to know what to focus on
when writing your paper. Consider the following general guidelines about how best to approach
writing a scientific paper:
1. Focus on argument. While it can be tempting to think about writing a paper as a means
to a grade, it is important to make a strong argument: focus on telling a convincing story about
science. Always start by analyzing your data, preparing any important exhibits, answering the
guiding questions, preparing a GRA, and outlining a detailed checklist for your paper.
2. Make appropriate claims. Don’t try to oversell your results. Make sure that the claims
you make are (a) supported by your data, (b) consistent with the design and objectives of your
experiment, and (c) reasonable in relation to similar, peer-reviewed results.
3. Don’t forget to make a GRA and a checklist. I’ve heard some rhetoric instructors
suggest that someone ‘just start writing, and it will come to you.’ This may work for some,
especially in some non-science disciplines, but it is a terrible idea in science. Starting by
sketching out your paper in a GRA and a checklist means that you are more likely to make a
convincing argument that follows a logical flow.
4. Engage with previous scientific studies and sources. A good argument is one that is
based on real scientific knowledge gleaned from reliable sources. As a result, claims that are
made must be supported by well-researched warrants (from primary sources) in addition to the
experimental evidence. Failure to cite these sources, in addition to being dishonest practice,
results in a poor foundation and weak scaffolding for your argument.
5. Each section stands alone. Experts rarely read papers from beginning to end. Frankly,
most experts start by looking at the exhibits (and the figure captions) and, in many instances,
48 · Undergraduate’s Guide to Writing Chemistry Papers
stop there. If an interested reader decides to “read the paper,” the likelihood is that they will not
start with the Introduction; rather, they will read the Results and Discussion and Conclusion
sections. The Experimental section is read when interested parties want more details about
how the results were achieved or by someone looking to repeat or extend the study. What this
means is that every section of a scholarly research paper needs to stand alone, without the need
for the reader to have read other sections. For instance, a Results and Discussion section always
starts by repeating in no more than one sentence the objective of the experiment.
6. Less is more, especially when it comes to data. Flooding your reader unnecessary data,
figures, and tables, can only lead to one thing: confusion. Only show appropriate exhibits that
need to be shown, and do it in the most appropriate way (Table, Figure, or prose). Before you
move on with your writing, start by making a list of the data that you have and make a plan
for how you will (or won’t) present it. Use GRAs and checklists to outline your work.
7. Have a good flow. Don’t write like you are checking boxes off a list (even though you are
checking off from your checklist). It can be easy to start each new component with a broad and
general declarative statement like ‘The results show ...’, ‘In conclusion we believe ...’, and ‘The
future directions are ...’. Instead, try to make sure that your paper has a natural and smooth
flow. Just make your claims, and discuss your results, without unnecessarily calling attention
to each component by name.
Writing Example 5.3: Good flow
Incorrect:
“The result of the manganometric determination of the % V
composition is that there is 45.
5
± 6.
6
% vanadium in the com-
plex. The theoretical value for the % V is 40%(*). The ex-
perimental value is statistically consistent with the theoretical
value. The relative standard deviation is 14.5%. The possible
source of error that led to the large relative uncertainty is like
the presence of bisulfite contamination.”
Correct:
“Managanometric titrations of compound 1 suggest a 45.
5
±
6.
6
% vanadium composition, which is consistent with the pre-
dicted formula of (NH
4
)
6
V
10
O
28
· 6H
2
O(*). The relatively large
deviation from the theoretical value (12.5%), as well as the
14.5% uncertainty, are expected with mangamometric titra-
tions of solutions containing small amounts of bisulfite.”
The second version relates the exact same information, but does so in a
concise manner that flows nicely.
8. Conventions of the discipline. Each section of the paper needs to conform to the
conventions that have been established by the scientific discipline in question. Always make
sure to consider the following details when writing your papers: (a) voice and tenses, (b) order
and arrangement of the paper, (c) the proper use of tables and figures, and (d) other genre-
Writing 5: Conventions of Scientific Writing · 49
specific restrictions. This is a veneer on top of the actual argument, but it is necessary in order
to get your work published and accepted.
9. Proofread. Though it should go without saying, you absolutely must proofread your work
to make sure that you’ve produced a concise, well-written paper that conveys a convincing
argument. Consider writing your paper early and allowing some time to pass from the initial
writing before you proofread – usually the next day works best. Also, it can be very helpful to
let a friend read it and not someone in the course. Your job is to be a teacher and presenter;
did they understand your argument?
10. Most importantly: less is more. This cannot be overemphasized: length has no corre-
lation with quality (though it can have an inverse relationship at times). On the contrary, overly
verbose rhetoric is unacceptable in scientific writing. Readers are not fooled by overly verbose,
unnecessarily long papers. The most successful papers are concise, direct, well-reasoned, sci-
entifically correct, and well-organized. Use GRAs and checklists to help prepare your writing
tasks and make the strongest argument possible.
5.3 Common grammar mistakes in (scientific) writing
Many people fall into the habit of writing as they speak. While this may be suitable for some
literary styles, it can often cause ambiguity in something meant to be read as technical writing.
Some of the most common mistakes seen from introductory (chemistry) students include poor
choice of words or phrases, wordiness, confusing word order, and poorly constructed modifiers.
Of course, keeping these things in mind when starting to write will help, but the easiest
way to catch the ones that slip through is to re-read your writing! Preferably, you should
not be reviewing just as you have written it, but rather after a break where you have distanced
yourself from the content re-reading immediately is significantly less effective because often
the context of what you were writing remains fresh in your mind and it is difficult to adequately
judge the writing (after all, you knew exactly what you wanted to say!).
5.3.1 Choice of words
Choosing the appropriate words is by no means a problem unique to science writing. Rather,
many introductory writing instructors often complain about some of the more confused words
found in student writings.
Below you will find a list of some of the most common mistakes in word choice amongst
novice writers. This list is by no means exhaustive, but should serve as an primer to the types
of grammatical errors that should be avoided.
Its vs. it’s: the often cited pet-peeve of writing instructors, these often confused words
are relatively easy to disambiguate. Its is possessive, while it’s is a contraction of ‘it is’
or ‘it has’ if you can’t replace it with one of these, then you are likely trying to use the
wrong one.
Their vs. they’re vs. there: Their is possessive (their piano is red). There is a place or
an idea (there were small amounts of white crystals that formed). They’re is a contraction
of ‘they are.’
50 · Undergraduate’s Guide to Writing Chemistry Papers
Ensure vs. insure: To ensure is to make certain that something occurs, while to insure
is to arrange for financial compensation in the event of loss or damage.
Then vs. than: Then is used to denote time, while than is used for comparison.
Proper word usage is of particular importance to scientific writing, because of the many
possibilities for confusion in meaning resulting from simple word choices. Many words have
connotations or implied meanings when used informally, but take on specific meanings in a
scientific setting. This can be a result of adopted colloquialisms and/or the use of gestures or
vocal inflection, which are lost when those same words are written, especially in a technical
context. Consider the following particularly common mistakes in novice science writing:
Data: more than any other word, data can be the most confusing to novice science writers.
Data is the plural form of datum, though the singular form is rarely used. Used correctly,
that would mean that you would write that “the data are” or “the data suggest,” rather
than “the data is.” That said, it is starting to become more common that the word data
is used in place of the collection of data, an uncountable collection of data. In this case,
we would write that “the [collection of] data suggests” and “the [collection of] data is”
(both in the singular). It is best to train yourself to use the proper, plural form of ‘data’.
Significant: when used in scientific writing, significant/significance implies a particular
outcome of statistical analysis (that your results are very unlikely to have occurred by
chance). Novice writers often speak of their significant results; they are trying to imply
that results are important, but the term significant implies statistical relevance. Instead,
the following words can be more appropriately used to convey your meaning: to indicate
something of a large effect (a big difference), use substantial; to imply the gravity or
importance of a result, simply use important; and to convey the gravity of a result, try
using serious.
Hypothesis vs. theory: a hypothesis is a proposed explanation made on the basis of
limited preliminary observations or evidence, usually as a starting point for further in-
vestigation, whereas a theory refers to a set of principles or system of ideas intended to
explain something.
Can’t vs. cannot: While not necessarily a source of confusion, contractions are dis-
couraged in scientific writing.
Efficacious, utilize, elucidate, and proximal: while none of these words are incorrect,
they are usually unnecessarily complex. When given a choice between a familiar and
a technical (or obscure) term, the more familiar term is preferable if it doesnt reduce
precision. Consider using the following as alternates: effective, use, explain, and close.
Additional examples of words and phrases that are often confused, or used incorrectly, are
tabulated in Table W5.1.
Writing 5: Conventions of Scientific Writing · 51
Table 5.1: Selected words or phrases that are often used incorrectly or in a confusing manner.
Category Example Explanation Sample Use
Comparison
Words
over/higher vs.
greater/more
than
Over and higher are less
accurate and also have spa-
tial associations
more than 50 mL was
needed instead of “over 50
mL was needed
fewer and less
fewer is used for num-
bered values, while less
should be used for quanti-
ties (except when units are
present then less is used
for the number)
fewer than 10 fractions
total versus less time
was needed (exception:
less than 10 mL”)
Confused
Words and
Phrases
affect/effect
Affect means to influence,
modify, or change. Ef-
fect, as a verb, means to
bring about, but as a noun,
means consequence, out-
come, or result.
The decrease in temper-
ature affects solubility.
The exposure to sunlight
effected the shelf life.
The difference in solvent
had a noticeable effect on
the rate.
whether or not
Only to be used to mean
“regardless of whether.”
Whether or not run
times are increased, the
accuracy will remain low.
NOT It was unclear
whether or not the run
times should be increased.
to comprise
Means “to contain” or “to
consist of”, NOT to be
used as a synonym for “to
compose.”
A molecule comprises
atoms. NEVER A
molecule is comprised of
atoms.
Subordinating
Conjunctions
while and since
vs. although/
because/
whereas
Although, because, and
whereas should be used
to join subordinate clauses
to the main sentence, be-
cause while and since
have strong connotations
of time.
Although the ini-
tial method was easily
perturbed by small vari-
ations in weight, the
final implementation was
considerably more robust.
52 · Undergraduate’s Guide to Writing Chemistry Papers
5.3.2 Words that aren’t actually words
For novices, knowing and using the proper term for something can often be challenging. Here
are a few of the more common made-up terms that creep into student writing.
Filtrated. Filtration is the process by which samples are filtered.
Massed. There is no verb for the process of determining the mass of a substance. Techni-
cally speaking, weighed is not the rigorously-correct term, since it is not the weight that
is being measured (though many people do use “weighed”). The good news is that it is
rarely necessary to mention that the mass of something was measured just report it!
Balanced. Chemical equations are balanced, but I think that students who write “bal-
anced” are referring to measuring the mass of something on a balance.
Put through. Unless you are writing about a needle and thread, this is probably inappro-
priate. Samples are not “put through” or “run through” a spectrometer. Spectral data
are recorded.
While this list is far from exhaustive, it gives you a sense for some of the more common mistakes.
For good ways to describe these things, look at the phrases used in your textbooks, lab manuals,
and the papers that you read.
5.3.3 Wordiness
In science writing, it is preferable to always be clear and direct. Superfluous descriptors and
excess (often redundant) words can confuse the reader and obscure your findings. Additionally,
there are page limits on most journal submissions, so being economical with your words is often
a necessity. Better to write with a clear and precise meaning as you go, rather than have to
come back and reword things later.
That said, sometimes longer phrases are used in good rhetoric. Use these longer phrases
for emphasis, and only sparingly. Consider the following examples of good places to reduce
wordiness:
a few instead of a small number of ”.
although instead of in spite of the fact that”.
because, since, or why instead of “the reason for, due to the fact that, in light of the fact
that, given the fact that, and considering the fact that.”
if instead of in the event that and under the circumstances in which.”
must or should instead of it is necessary that and cannot be avoided.”
5.3.4 Sentence construction and word order
Here are some of the more common ways in which sentences are improperly constructed:
Writing 5: Conventions of Scientific Writing · 53
Ambiguous pronouns. When you use a pronoun, make sure that the noun to which
the pronoun refers is clear. Ambiguous pronouns are most often seen when using this
and that. Be sure to include the noun if there is a possibility of confusion.
Writing Example 5.4: Ambiguous pronouns
Incorrect:
The reaction was complete when an orange precipitate formed
and it was collected via filtration.
Correct:
The precipitate was collected by filtration and provided 417 mg
of the orange product with 53% yield.
It is not clear in the first sentence to which noun ‘it’ is referring.
Too many short declarative sentences. Although they are often the most clear to
read (and easiest to write) you do not want your writing to suffer from poor flow and
cause a disconnect between ideas. Try combining very short sentences to help related
ideas move smoothly.
Double Negatives. While double negatives are sometimes used as a stylistic nuance to
indicate an understated affirmation, they are considered improper in scientific writing.
Writing Example 5.5: Double Negatives
Incorrect:
The difference in absorbance values was not unexpected.
Correct:
We knew it was possible for the absorbance values to differ.
Use affirmative language rather than a double negative.
Misplaced modifiers.
“Based on” vs. “on the basis of”: Based on is used to modify a noun or a
pronoun (often immediately preceding or following), while on the basis of is used to
modify a verb. Think of it as something can be based on something else, whereas
you would do something on the basis of something.
“Due to”: This phrase is used to modify a noun or pronoun directly before it in
the sentence or following a form of the verb “to be.”
Split Infinitives: In non-technical writing split infinitives are accepted as gram-
matically correct this is not the case in scientific writing. A split infinitive allows
a statement to be interpreted in multiple ways. In science writing this varied inter-
pretation is what you are trying to prevent by choosing the words carefully.
54 · Undergraduate’s Guide to Writing Chemistry Papers
Writing Example 5.6: Spit infinitives
Incorrect:
Each data point was carefully measured after the initial sample
preparation.
Correct:
Each data point was measured after the initial sample prepa-
ration.
Removing the adverb ‘carefully’ removes the split infinitive.
Restrictive Expressions. If a phrase or clause is necessary for the sentence to
make sense it is restrictive. Restrictive clauses should be introduced by that not
which. A phrase or clause is considered nonrestrictive if the information added is
not essential. Nonrestrictive clauses and phrases are set off by commas and can be
introduced using which (or who). A good way to tell the difference is to check that
if you deleted the clause/phrase, that the sentence would still make the intended
point to the reader or relate the desired information?
Writing Example 5.7: Restrictive Expressions
Incorrect:
“Before titration could begin, the flask was lowered to a
temperature which would prevent evaporation.”
“Plotting all four of the UV/Vis curves that are shown in Figure
3 allowed us to select the optimum starting concentration.”
Correct:
“Before titration could begin the flask was lowered to a
temperature that would prevent evaporation.”
“Plotting all four of the UV/Vis curves (Figure 3) allowed us
to select the optimum starting concentration.”
The first example is clearly restrictive because if the clause was deleted it
would not convey the importance in lowering the temperature, or to what
sort of temperature it was lowered. In the second example the reference
to the figure is made concisely.
Commas. While in the United States it has become common practice to omit the
comma before the word ‘and’ in a list of three items (this is called the “Oxford comma”).
In fact, there are some that argue that the Oxford comma is actually incorrect in the
US. Unfortunately, the absence of the serial comma in a list of three items can lead
to substantial ambiguity. In science writing, as it is in all writing in the rest of the
English-speaking world, the Oxford comma is mandatory.
Writing 5: Conventions of Scientific Writing · 55
Writing Example 5.8: Commas
Incorrect:
For breakfast today, I ate eggs, toast and juice.
Cobalt complexes are important, industrial compounds.
Correct:
For breakfast today, I ate eggs, toast, and juice.
Cobalt complexes are important industrial compounds.
The first sentence was incorrect because it was missing the serial (or “Ox-
ford”) comma the ambiguity in the first version is that it is possible to
imagine that the juice was on the toast, rather than with it. The second
sentence was incorrect because commas can only be inserted between two
or more adjectives if the adjective order can be reversed without losing
meaning.
5.4 Lessons from experts in the field
There is no substitute for the experience that comes with writing, and reading, scientific
papers over the course of a number of years. Three of the more important lessons are (a) that
you need to know how people read scientific papers before you can write a good one, (b) that
each section of a paper will stand on its own there is no need to read through a whole paper,
and (c) that experts use an array of tools to assist in their process.
5.4.1 Reading scientific papers
Part of being a research-active scientist is keeping-up with advances in the field by reading the
scientific literature. Rarely will someone who is browsing the scientific literature read an entire
paper from start to finish. Rather, we tend to start by reading the Abstract (to be discussed in
a later chapter). After the Abstract, the next stop is the conclusion section and a look at the
pictures (figures, tables, etc.) the remainder of the paper would only be read if those seem
interesting.
5.4.2 Each section stands alone
As we’ve mentioned previously, people are usually reading a paper for a very specific purpose: to
extract some necessary data (Results), to glean some understanding of an important/interesting
phenomenon (Discussion), to learn how to perform a specific procedure/synthesis (Experimen-
tal), or to keep on top of the newest innovations in the field (Abstracts, Figures, Conclusion).
Notice that none of those reasons require the reader to agonize over the entire paper; rather,
we will often only read the section(s) of the paper that are necessary for our purposes.
Consequently, it is important that each section be able to stand alone. The Introduction sec-
tion will always contain the motivation and objective of an experiment, as well as a description
of the general approach taken to solve the scientific problem. The Conclusion section begins
by summarizing the overall success/failure of the experiment in the context of the motivation.
56 · Undergraduate’s Guide to Writing Chemistry Papers
All that remains is to provide a brief and concise summary of the problem to be solved at the
beginning of the Results and Discussion. The captions of tables and figures (or a rephrasing of
them) appear in the text as the table/figure is being discussed.
5.4.3 Use tools to effectively plan your writing
Experts develop effective tools and devices to help them in their workflow and process. For
expert science writers, many of these tools are quite “low-tech” and are easily accessible for
novices entering into the field. Consider adopting the use of some of these to help you in your
work.
Voice recorders or voice recording applications can be very helpful for organizing your
thoughts and getting the major points down. Talking out loud, as if to another person,
is often the best way to capture ideas as you have them and to refine them. Later, once
you’ve articulated how you want to say something, you can extract the phraseology from
the recording. While this may not save you a lot of time (though it can), it definitely
reduces the frustration of losing your thoughts or that “perfect way” you thought to say
something.
Reference managers are very versatile and powerful ways to keep track of the research
you do in the literature. Many of them can keep the actual PDF of the papers you read,
manage the annotations and notes you make, and allow you to classify papers by topics
and more. While some of these packages can be costly, many free (and yet powerful)
options are available.
A good printer, or a cheap printer, will allow you to print drafts and partial drafts for
editing ease. If you find it cumbersome to edit your work on a screen, try getting an
inexpensive printer to print drafts.
GRAs and checklists, while not technology, are two of the most important tools that you
can use. Organizing and solidifying your thoughts, before starting to write, is the best
way to streamline the writing process.
5.4.4 Don’t try to sound technical, and know your audience
Let the science wow your audience, not your verbiage. Unfortunately, many students new to
the sciences conflate the difficulty of a subject-matter with the complexity in the writing; these
students tend to use overly complicated words and language, such as efficacious and utilize,
instead of more appropriate words like effective and use. Equally problematic is the use of
flowery and exaggerated language. Remember, the purpose of a peer-reviewed paper (Journal
Article) is to communicate new findings, discoveries, and theories to the greater scientific public;
to that end, always make sure to be as clear and concise as possible in your scientific writing.
That said, don’t confuse brevity, simplicity, and concision for lack of detail. In other words,
don’t be vague. Make sure that your arguments are specific and that you use enough detail
to make your point. You have to know your audience. It would be completely inappropriate
to spend any time in your paper trying to convince your reader that Beer’s law is meaningful;
if they don’t at least know that, then they have no business reading your paper
3
. Knowing
3
We’re assuming that we are writing to an expert, or at least scientific, audience.
Writing 5: Conventions of Scientific Writing · 57
where to strike the balance of brevity with sufficient detail is a hard thing to gauge. Many
career-scientists struggle with these same issues.
5.4.5 Be straightforward about the limitations of your work
One of the hardest things for students to do is to resist the compulsion to rationalize, or ignore,
‘bad’ data. The results that you obtain in any given experiment are, for better or for worse, what
you have to work with. Please refrain from claiming that an unreasonable difference between
your results and your expected results is anything other than what it is: an undesirable outcome.
Some people feel that they need to justify getting bad results in order to be persuasive.
This is a mistake. Instead of claiming that a bad result is acceptable, consistent, etc., you
should simply present an argument as to why you may have received the erroneous results.
Additionally, reports of bad results would never actually get published in the literature. Nor
would anyone try to publish them. This is another distinction between classroom and real-world
writing.
5.4.6 Putting it all together
Although the mistakes and tips discussed in this chapter will probably not often be explicitly
looked for in grading considerations, their proper implementation will only serve to improve
the quality of your writing (and your grade). If a grader is able to easily read your paper and
find the relevant information, there is no risk of lost points due to simple lack of clarity (not
to mention a reader who isn’t frustrated by confusing sentences and conflicting information
will probably be a little more generous in their evaluation!). Writing well also helps you as a
scientist, since you must thoroughly think about how concepts relate or what results mean in
order to create logical, smooth flowing sentences about them. If you haven’t thought about
what a result means in the context of the experiment you just ran, it is very difficult to convey
it to someone else.
5.5 End-of-chapter assignment
1. Use the information in Section W5.1 to critically analyze each of the following excerpts
from actual student papers. For each, indicate what general guidelines have been ignored
or violated, and propose a version that is more correct (it is possible that more than one
thing is wrong with each). Note: where you feel that a reference is merited, simply indicate
with a superscript. There is no need to actually look up references for this assignment.
(a) The experiment is designed to introduce the student to the industrial chemistry
process, and give an idea of how the basics work.
(b) While this objective was trivialized with the use of a spectrophotometer, the infor-
mation gained from acquiring the absorbance of the dyes is the true matter being put
to investigation. By putting the dyes through a spectrophotometer, the absorbance
and wavelength of the dye was established. Why was knowing this important? The
true objective of this lab was to test the particle-in-a-box model, and test it for its
accuracy in predicting the wavelength of light emitted from the dyes.
58 · Undergraduate’s Guide to Writing Chemistry Papers
(c) With the accuracy of determining Avogadro’s number hinging upon multiple factors,
it is to no wonder that advanced researchers have shifted away from measuring N
A
in such error-ridden experiments and towards more exact methods such as X-ray
crystallography.
(d) Any other error was either systematic or human, and would account for faulty data
and inconclusive results.
(e) The relative standard deviation was also 5.45% showing that the lab was executed
with some precision as well.
2. Identify the grammar and writing problem(s) with each of the following writing samples
and propose a correct version.
(a) The resulting mixture is a heterogenous solution, but is mainly comprised of hexane.
(b) Because of the compound’s sensitivity to light, it is not uncommon to see a slight
decay pattern over time that is depicted in Figure 5.
(c) We weren’t able to determine the identity of the complex strictly by color, since
there was more than one ion complex that could have been green.
(d) The uncertainty introduced into the experiment due to wet glassware probably had
the large effect on the final results, which is seen the large percent difference from
the literature value.
(e) The average of the three trials from the UV-Vis method was 6.9832± 0.004mg. The
Average of the three trials in the AAS method was 7.82± 0.45mg. One point from
each method had to be dropped because it failed the Grubbs test, meaning it was
an outlier relative to the rest of the data. A Student’s T test was also performed to
compare the results of the data of both methods. Both methods were compared to
the literature value of 7.0mg/L.
8
The average for the UV-Vis method had a value
less than t
table
, meaning that are statistically similar. The AAS method had a value
that was greater than t
table
, meaning that it failed.
3. Give a copy of your last Abridged report to a friend in your class (someone that you
trust and who you respect) and get a PDF copy of their paper. For each of the following
questions, please mark up the PDF directly.
(a) Identify two things (sentences, paragraphs, etc.) that you feel they did especially
well. What about them do you appreciate?
(b) Identify two places where you feel they could have made better grammatical or word
choices. Suggest a better version for each.
(c) Find and list two things that you would improve about their paper. Do not include
places where you feel they could improve their grammar or word choice; rather focus
on the organization, presentation, and content of their paper. Explain each briefly
(in no more than one sentence).
(d) Do you feel that they have made a convincing argument? Explain briefly (no more
than a couple of sentences).
Writing 5: Conventions of Scientific Writing · 59
(e) Do you feel that they have written a concise paper that follows a logical flow? Do
you feel that it is pitched towards an appropriate audience? Explain briefly (no more
than a few sentences).
4. Preparing an abridged paper for a recent experiment:
(a) After completing your data analysis, prepare a graphical representation of argument
(GRA) to help you plan your paper. Make sure to do enough research to ensure that
your answers are based on good understanding of the material.
(b) Prepare a checklist that you will use to write your abridged paper (you should not
include Introduction and Experimental sections).
(c) Based on your GRA and your checklist, use the information in this chapter (and all
of the previous chapter) to write an abridged paper (Title, Results and Discussion,
Conclusion, References, and Supporting Information) for your most recent lab. Make
sure to include your GRA and checklist.
Reminder about papers submitted
The paper always begins a new page (with a title), followed by the References (these do not
start on a new page), and then the Supporting Information (starting a new page, after the
References). Note: you will write the Supporting Information before you write the paper, but
it will appear at the end.
60 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 6: RESEARCH, SCIENTIFIC LITERATURE, AND
ANNOTATED BIBLIOGRAPHIES
In this chapter we will focus on four major goals: (1) methods for purposeful reading of papers,
(2) strategies for finding relevant sources in your research, (3) referencing the works that you
used, and (4) preparing meaningful and useful annotated bibliographies from your research.
6.1 How to read scientific papers
There are a number of good reasons why you would likely be reading a scientific paper. Amongst
them are: gathering specific data, researching the current state-of-the-art in a field, preparing
for a ‘journal club,’ and keeping up with the literature in your field.
How you approach reading a paper is very dependent on what you hope to accomplish a
partial list is presented in Table 6.1.
Table 6.1: Common goals for reading scientific papers
Goal Primarily-relevant sections
Learning new method / approach Methods, Introduction
Understanding limitations of a method Conclusion, Discussion
Significance of a research area Introduction, Conclusion
Looking up data / values Results, Figures
Broad interest in the field Abstract, Conclusion
To help illustrate, let’s examine a couple of approaches that experts have found to be helpful
when reading research papers. Using either of these methods, you should be able to extract
the gist of the argument from a scientific paper using a quick scan only a few minutes. In
order to learn the material well, you will likely need to spend more time working through the
authors’ data and claims, and taking substantial notes.
You will find it useful to make detailed annotations on the papers that you are reading.
Additionally, many digital reference managers will allow you to save an annotated copy of the
paper. Most of these programs can even assign keywords to the papers so that you can easily
search for the concepts that you want to find at a later time. Keep in mind, when scientists
say “read a paper,” they really mean digest and annotate a paper.
61
62 · Undergraduate’s Guide to Writing Chemistry Papers
6.1.1 Sample approach #1
First, take a look at the schemes/figures that the authors have chosen to show within the
paper. Ask yourself: (a) What reactions are being shown in the schemes? At first glance, do
you understand these schemes? (b) What structures are they choosing to show? Why?
Next, try to very generally interpret what the authors are showing in the figures in the paper
(e.g., kinetic data from UV-vis, redox potentials from electrochemistry, molecular structures
from x-ray crystallography, etc.) This can usually be done by inspecting the axes and the
captions.
Third, read the conclusion of the paper. The conclusion is a very short and succinct section
that conveys the overall findings of the article and, typically, what the authors believe is most
significant.
Finally, read the abstract (if there is one), because this gives a general overview of the
findings in the paper. If the goal is to understand the paper, then you are likely done. If,
however, your goal is to present the paper to others, then read through the paper fully (except
for the methods) and take notes on each area of interest.
The first approach is a good way to approach working through a paper that you already
know will be helpful. If, however, you are not sure which articles you’ve found will be useful,
you will find it useful to try another approach. Specifically, use abstracts (skim them) to narrow
down the articles that sound relevant. The abstract should summarize their main objectives.
Next, follow the steps in approach 2 to work through and make sure the paper is appropriate.
6.1.2 Sample approach #2
Start by skimming through the figures and read the captions to make sure their data match
their objectives indicated in the abstract. At this point, try to interpret the data without
authors’ narrative (i.e., don’t read their conclusions).
Look at the bold headings in the methods section to determine whether or not you are
familiar with their techniques. If there are techniques that you don’t recognize, use a search
engine to gather background information.
Read the last paragraph or two of the introduction to learn about their motivation, objec-
tives, and findings (argument summary). For in-depth understanding of the paper (i.e., journal
club presentation), read through the results and discussion section and go back to the figures
to confirm authors’ arguments.
The conclusion section is read last and the methods section need only be read if you want
to reproduce their work.
6.2 Tools and strategies for finding relevant papers
6.2.1 Searching for references common strategies
There are several strategies one can use to perform an effective search of the literature. The
type of search will depend on what type of information you need. For example, it might be
quicker and easier to perform a regular Google search for vitamin C to find out its structure
than to spend time logging into SciFinder Scholar. However, if you wanted to know who sells
it for the cheapest price, you might use a much different strategy. You will find below several
strategies for making the most of your literature search.
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 63
Keyword/Topic Search: The most common search for beginning scientists is the key-
word or topic search. Because you are unfamiliar with the literature, this can also be the
most difficult search to perform. There are two main ways to search by topic, each with
their own merits. You may find that by using both strategies, you obtain a more complete
set of references than by just using one technique:
Start general and become more specific if your goal is to find introductory-level
information on a topic, it is best to start with general search terms. You can then
use the filtering capabilities of each program to narrow the large list of references
into something more manageable (ie., journal name, author, year, document type,
institution, etc.). Refining by topic is also a good way of becoming more specific.
Start specific and become more general
if you are looking for information that is
specific to a particularly narrow area of research, start by using search terms that de-
scribe exactly what you are looking for. If you are unable to generate any hits, make
your search terms more general by eliminating words, using broader terminology,
and not using any filters.
Cited/Citing Reference Search: SciFinder Scholar, Web of Knowledge, and Google
Scholar (more on these tools later) are all good choices for searching for a particular
article, references it cites, and references that cite it. However, the databases each program
searches are different, so you may find slightly different results. Simply enter the article
citation information into each program. Once you find the reference, click on “Get Cited”
or “Get Citing” in SciFinder Scholar (similar terms in Web of Knowledge and Google
Scholar) to generate a new list of references which you can peruse. Another useful feature
of Google Scholar is the link “Related articles” which will lead you to tangentially-related
citations.
Author Search: All three search tools are also good choices for searching by author.
Simply enter the author name to see which references have been published by a partic-
ular author. Be cautious of authors with the same name (this is easy to catch if you
are searching within analytical chemistry and retrieve additional hits about evolutionary
biology). Also be aware that some publications might have different versions of the same
person’s name (e.g., J. Harden vs. J. D. Harden vs James D. Harden). SciFinder offers
the option to display all results of name variations if they show up.
Structure Search: Using just those programs introduced to you in this chapter, search-
ing by structure is only possible using SciFinder Scholar. Its user-friendly and intuitive
chemical drawing program will enable you to search for chemicals that have been indexed
by Chemical Abstracts. Each result is linked to reactions containing it and references for
its use, synthesis, or general study. Simply choose “Get Reactions” or “Get References”
to explore.
Combining Search Tools: SciFinder Scholar, Web of Knowledge, and Google Scholar
each have their own set of strengths and weaknesses. Knowing when to use each of
them comes with practice but is important to obtaining the information you need in
a reasonable amount of time. Fortunately, the information you glean from one search
engine can be used to perform a search in another. SciFinder is really good at finding
64 · Undergraduate’s Guide to Writing Chemistry Papers
references on a research topic. Once found, you may search Web of Knowledge for its
cited/citing references or search Google Scholar for related articles. Likewise, Web of
Knowledge is excellent at performing author searches. But if you wanted to know if a
particular author has ever published a synthesis of a particular compound, you would
need to follow up with a SciFinder Scholar structure search. Thus, there are many ways
to “mix-and-match” search tools to gain maximum coverage of the literature.
Other Helpful Hints: Boolean operators such as AND, OR, and NOT help you filter
search results from the beginning and limit the generation of false or misleading hits
(all capitols must be used). Wildcards are special characters (such as *) that permit
alternative spellings or suffixes to your search terms (for example, the term electrochem*
will return hits that include electrochemistry and electrochemical). In addition, placing
a term in quotation marks will ensure that a particular word or phrase appears in your
results set exactly as written.
6.2.2 What references are acceptable?
We’ve already discussed what needs a reference: anything that is not common knowledge and,
therefore, required research. But, how do you know which references are good and which ones
are not? To be frank, this comes with practice, but you can become better at it by reading the
primary literature.
For beginners, there are several criteria that will help you decide which references are ac-
ceptable:
Type of Source: For the introduction section of your write-up, books and review articles
are the best sources of general information on a research topic or experimental method.
You may find these citations by refining your search in one of the databases (discussed
below) by reference type. Other sections of your write-up may require you to cite a specific
article and it is standard practice to find more recent papers that cite this lead article
as a reference - someone may have recently expanded upon this work. internet references
are rarely acceptable.
References within a Source: An addition supply of leading references may come from
the introduction section of a research article on a topic. If the article is recent, then
someone has already done the literature searching for you! It is good practice to look up
all of the references found in the introduction of a journal article to better understand a
topic.
Timeframe: The most current information is probably the best indicator of the state-
of-the-art, so you should choose references that have been published in the last 15 years.
However, a seminal publication on a topic, often published decades ago, should also be
referenced to highlight major contributions to the field. Thus, there is a balance between
being up-to-date and respecting the scientific historical record.
6.2.3 What references are not acceptable?
There are three major things to avoid in choosing your references: internet references (see
below), foreign-language references (we will restrict our sources to English only), and the ab-
stracts/proceedings of meetings or conferences. The major reason to avoid internet references
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 65
is because of their reliability. Foreign journals and the proceedings/abstracts of meetings are
certainly reliable, but you couldn’t have possibly used them to glean information. After all,
you didn’t attend that conference!
An important note about the use of electronic references
There is a difference between an “internet reference” and a reference
accessed via the internet. Primary literature that is accessed via the
internet should NOT be cited as an internet source. If journal articles,
books, and similar documents, are accessed online through the databases,
they are references like their corresponding print forms.
In addition, a common misconception regarding the use of material on
the internet is that it is by its nature “free.” Despite “open access”
to the internet, someone owns a copyright for information on a web-
page and thus permission must be granted from the copyright holder
for reproduction (including all pictures, diagrams, photographs, etc.). It
is standard practice to reference these sources. Failure to do so is plagiarism.
Also, remember that sites such as Wikipedia, About.com, and even search
results from Google can be scientifically unreliable due to a lack of oversight.
These are not peer-reviewed sources of information and are not necessarily
written by experts. While certain facts may be correct (molecular weights,
densities, etc.), it is best to check the primary literature for confirmation
(often these sites have references at the bottom of the page, which you can
follow and read for yourself). Any interpretations or conclusions regarding
data should not be referenced.
Finally, work that is published online by another student may not be used
under any circumstance, nor may it be referenced.
6.2.4 Using the search tools to find references
At this point, you may be asking yourself “How do I find the references I need?” Fortunately,
there are several tools readily available that will help you search the literature. While each of
them is designed slightly differently, the same general themes apply no matter which one you
are using.
CRC Handbook of Chemistry and Physics
The CRC Handbook of Chemistry and Physics (often referred to only as the “CRC Handbook”)
is a comprehensive reference work that comprises literally hundreds of tables of scientific data.
Included in these tables are physical constants, chemical properties, mathematical data and
functions, and laboratory safety information. This is a really good place to start when you are
searching for values, constants, and properties that you have reason to believe are well known.
66 · Undergraduate’s Guide to Writing Chemistry Papers
Most libraries will have a physical copy of the book, but the online version allows user to
search by chemical structure (written or drawn), which makes it even more powerful and easy
to use. Most research universities have site license access to the CRC Handbook, which can be
accessed at http://hbcponline.com/.
Note: although this a website, it is a web representation of a physical book (like most only
references and journals). The 2017 CRC Handbook is cited as: CRC Handbook of Chemistry
and Physics, 98th ed; Rumble, J. R., Ed.; CRC Press: Boca Raton, FL, 2017.”
SciFinder Scholar
SciFinder Scholar is the most widely used scholarly database in chemistry. It contains entries
from two major indices: Chemical Abstracts and MEDLINE. Chemical Abstracts indexes new
compounds and articles that appear in nearly every type of print media (book, journal, thesis,
technical report, etc.). MEDLINE is a database of published papers related to medicine and
biology. Together, they cover most of what chemists need. Note: you will need to be connected
to the BU network, or have the VPN active (http://goo.gl/BRYFqm), to register for SciFinder.
To access SciFinder, follow these steps:
1. Go to the following URL: http://goo.gl/KpL2v and then follow the instructions to create
a new account. You will receive an email containing a link for activating your account.
This link will expire within 48 hours and you will not be able to register twice, so be sure
to activate your account right away.
2. Once your account has been activated, go back to the same URL and click “Connect
to SciFinder Web.” It is recommended that you bookmark this site. Using your newly
created username and password, login to SciFinder. Note that the university has only
a limited number of licenses that can be used simultaneously, so if you receive an error
message indicating this, try again in a few minutes.
3. To search the database, you have several options. At the top of the page are three icons
that will serve as the main criteria in your search. Sub-criteria are found on the left-hand
side of the screen.
Other useful features of SciFinder Scholar include:
The “Tools” menu allows you to further refine your answer set. You may remove dupli-
cates, combine answer sets, etc.
The “Get Related” menu allows you to view all of the references that are cited in an
article, as well as all of reference that have cited that paper since.
The “Analysis” and “Refine” options also allow you to categorize your results based on
criteria you choose.
Finally, you may save your search results for the next time you login.
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 67
Web of Knowledge
Web of Knowledge is an interface that allows simultaneous searches of several databases: Web
of Science, Biological Abstracts, MEDLINE, and Journal Citation Reports. Web of Knowledge
often contains information that is unavailable in SciFinder and the search engine works similarly
to SciFinder, so it is another reliable resource for finding appropriate references. To access Web
of Knowledge, follow these steps:
1. Go to the following URL and login to authenticate: http://goo.gl/vgFtF. It is recom-
mended that you bookmark this site.
2. You may begin to search the database by topic, title, author name, etc. You may also
search these criteria simultaneously by adding another field.
3. Refining your answer set is made possible by choosing from one of the tabs on the left
hand side. You may also find the cited and citing references.
4. Clicking on the “FIND BU” link will allow you to directly access the full text.
Google Scholar
The most popular search engine for non-scientists is Google. Unfortunately, searching for sci-
entific information on Google almost always directs you to information that does not meet the
peer-review standards of the scientific community. For example, a Google search for electro-
chemistry of batteries will direct you to a Wikipedia page on the electrochemical cell hardly
the kind of literature you need for scientific writing. Luckily, Google Scholar is a search engine
that is designed for the academic world. It works just like any regular open-access search en-
gine except that is searches the primary literature for your topic of interest. To access Google
Scholar, follow these easy steps:
1. Go to the following URL: http://scholar.google.com. Note, this can also be accessed
via the main Google homepage. It is recommended that you bookmark this site.
2. Search for a topic of interest directly or click on “Advanced Scholar Search” to provide
more details.
3. Within “Advanced Scholar Search,” you have the option of searching by topic, author,
publication name, etc.
4. You may include/exclude patent literature and limit your search to a particular subject.
5. If you are on a BU network or have VPN active the “Find Online at BU” link may appear
on the right side of the search result. These links appear only for those results that are in
journals for which BU has a subscription, and will take you to the full text, if available.
6.3 Working with references in your scientific papers
Journal articles, letters, and research proposals all contain a References section that lists articles
from the primary and secondary literature that were used in the preparation of new work. A
proper References section includes scientific sources that are relevant, respected, and current;
more importantly, it contains the actual sources used in the research for the paper.
68 · Undergraduate’s Guide to Writing Chemistry Papers
6.3.1 Why is citing the literature necessary?
All scientific progress, in general, is an extension of work done previously. Early scientists
documented and kept track of each other’s contributions. When any new discovery was made,
they responded by “referring” to what was already known about a particular topic and where,
when, and by whom it was published. Today, we continue this practice by appending a Ref-
erences section to our own work. In this section, we cite a piece of original research or major
contribution to the field. By “referring” to a body of work, we acknowledge pioneering work in
a field, assign credit for a particular intellectual discovery to an individual or group, supply a
source of additional reading on a topic, provide a historical perspective and/or context to any
new work, and ensure high ethical standards in science.
Using references also allows you to keep your writing more concise, that way a reader
who is familiar with a topic isn’t burdened with detailed explanations of background that
should be well-known, but someone who isn’t well versed knows exactly where to go to find the
information.
Understanding the contributions that others have made to the field is critically important as
a young scientist. As you begin to read the scientific literature, pay close attention to not only
the content of the text, but also to the list of references included therein. After all, a common
complaint in peer review is an inadequate understanding of the scientific literature and could
cost you publication.
6.3.2 What needs a reference?
We will explain how to format references within a text and in the References section itself later,
but first you may be wondering when to use references and how frequently they are needed. In
general, when you are writing any type of scientific communication, references must appear for
any idea that you cannot claim as uniquely your own. That means that you need to provide
references for things like:
Procedures: that you followed, or even borrowed from, in the process of doing your work.
For example, students following a procedure from a guide (like a lab manual or journal)
for doing iron-content analysis would need to reference the guide.
Concepts: especially when it comes to novices, there will be a lot of references to books,
articles, and reviews, that explain fundamental concepts that are needed for understanding
the work (both its motivation, the methods used, and the outcomes).
Other studies: science doesn’t happen in a vacuum we will often be doing work to
extend, refute, support, or refine previous work. Those studies must be cited.
While citations should appear for any source that has been influential for the present study,
an exhaustive list of references is not always necessary. For instance, suppose that you are
reading a journal article that references three others; even though all four sources were used,
you may choose to only cite the one, lead reference.
Finally, information that is commonly accepted knowledge does not require a reference.
It can be difficult to know where to draw the line when it comes to this supposed ‘common’
knowledge. For now, let us assume that anything that you needed to research is not common.
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 69
References in a Journal Article
Since most of this guide is devoted to learning about writing journal articles, let’s investigate
the nature of references by section of the journal article.
Title and Abstract: no references appear in these sections.
Introduction: the introduction of a scientific paper includes the motivation for the study
and a thorough review of the state of the art of the field (i.e. background information).
Both of these areas will involve substantial references to the literature.
Results and Discussion: while the results of your experiment will likely not require
citation, the discussion will usually be very heavily cited. Make sure to include all of the
sources that you referenced when analyzing and interpreting your results, the sources of all
previously-referenced results and values, and anything you used for drawing conclusions.
Strong arguments will be those that are rooted in thorough research.
Conclusion: citations are common when discussing any future directions, or extensions,
to the work. Additionally, the global ramifications of your work are usually based on
research as well.
6.3.3 Format of citations in the body of the paper
Citing references within the text will largely depend on the editorial conventions of the particular
publication. Most of the chemistry journals, as well as the papers that you write a student in
a chemistry course, follow the standards from the flagship chemistry journal of the American
Chemical Society (The Journal of the American Chemical Society, commonly referred to as
JACS). The complete details of the style and formatting can be found in the ACS Style Guide,
but here are the most important rules to follow when using references within the text:
References are numbered chronologically starting at the beginning of the document, in-
cluding those in tables, figures, and graphs. Don’t forget to check your references if you
switch sentences around. When a source is cited more than once, the numbering follows
its first instance.
Citations in chemistry use Arabic numerals (1, 2, 3) only.
A superscript should appear immediately after an idea requiring a reference has been
first introduced. If the reference applies to an entire sentence/clause, the superscript will
appear after punctuation.
Example: Superscripts for citations
Phenolphthalein has been used extensively as a pH indicator in titrations.
1
Author last names may appear in the text when referring to particularly noteworthy work.
If a reference has two authors, both authors’ last names should appear separated by the
word and. If a reference has more than two authors, the corresponding author (usually
the principal investigator) should be listed, followed by the phrases et al., and coworkers,
or and colleagues.
70 · Undergraduate’s Guide to Writing Chemistry Papers
Example: Referencing authors by name
A working electrode was made from platinum as reported by Haddon and
Loria.
2
A reference electrode was constructed from Ag/AgCl as described by
Jenkins and coworkers.
3
If you cite two references for a single idea or item, superscripts should appear consecutively
with a comma, but without a space. If more than two references are cited, a range should
appear (ascending and using an en dash).
Example: Multiple references
Palladium-mediated reactions are particularly important for C-N bond
formation.
4,5
Olefin metathesis catalysts have revolutionized pharmaceuti-
cal C-C bond formation.
68,10
A reference should be cited at a reasonable point in a sentence or clause.
Example: Location of citation
... recent studies suggest
4
... was reported
1,35,7
... as described
1416
...
previous results indicate.
3,4
6.3.4 Format of the References section of your paper
The references section, located immediately following the Conclusion section of a paper, is
where the information on all the references are listed that correspond to in-text citations. This
makes it easier to find references on a particular topic or from a particular author since only the
most important information is included in a citation. Just as mentioned above for the in-text
citations, you will use the JACS formatting criteria. Here, we will go through the conventions
that you will apply to your own references section.
A References section at the end of a document should list the references in chronological
order as they appear in the text, with numbers. Depending on the source, there are a few key
criteria that are required to ensure proper referencing. The following list shows you how to
properly cite a particular source in the References Section:
Book (no editor): Author 1; Author 2; Author 3; Book Title, Edition Number (if
any); Series Information (if any); Publisher: Place of Publication, Year; Volume Number,
Pagination
Example: Referencing a book (no editor)
Grossman, R. B.; The Art of Writing Reasonable Organic Reaction Mecha-
nisms, 2
nd
Ed.; Springer-Verlag New York, Inc.: New York, 2003; pp 23-25.
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 71
Book (with editor): Author 1; Author 2; etc. Chapter Title. In Book Title, Edition
Number; Editor 1, Editor 2, etc., Eds. Series Information (if any); Publisher: Place of
Publication, Year; Volume Number, Pagination.
Example: Referencing a book (with editor)
Paolesse, R.; Monti, D.; Dini, F.; Di Natale, C. Fluorescence Based Sensor
Arrays. In Topics in Current Chemistry: Luminescence Applied in Sensor
Science; Prodi, L; Montalti, M; Zaccheroni, N, Eds. Springer-Verlag: Hei-
delberg, 2011; 301, 139-174.
Journal article: Author 1; Author 2; etc. Journal Abbreviation, Year, Volume (Issue,
if any), Inclusive Pagination.
Example: Referencing a journal article
Merrifield, R. B. J. Am. Chem. Soc. 1963, 85 (14), 2149-2154.
Note: list all authors (if fewer than 15). Use the appropriate journal abbreviation from
the Chemical Abstracts Service Source Index (CASSI). You can search for the appropriate
abbreviation at the following link: http://goo.gl/q44au. A brief list can also be found
in the ACS Style Guide.
Patent: Patent Owner 1; Patent Owner 2; etc. Title of Patent. Patent Number, Date.
Example: Referencing a patent
Brewer, A. K.; Knight, J. Pyrrolopyrimidines as Antineoplastic Agents.
WO/2011/095246, 2011.
Thesis: Author. Title of Thesis. Level of Thesis, Degree-Granting University, Location
of University, Date of Completion.
Example: Referencing a thesis
Boone, A. C. PNP Ligands in Asymmetric Catalysis. Ph.D. Thesis, Univer-
sity of Notre Dame, South Bend, IN, 2004.
Electronic Reference: Author (if any). Title of Site. URL (accessed Month Day, Year),
other identifying information (if any).
Example: Referencing the internet
Sigma-Aldrich Home Page. http://www.sigmaaldrich.com/united-
states.html (accessed July 13, 2011).
72 · Undergraduate’s Guide to Writing Chemistry Papers
Mentors, Instructors: Conversations with Name 1, School, City.
Example: Referencing Instructors
Conversations with Dr. John Doe, Boston University, Boston, MA.
Peers, Classmates: Conversations with Name 1, Name 2, and Name 3 (Course Code -
Course Name, Year), School, City.
Example: Referencing classmates
Conversations with John Doe and Jane Smith (CH109 P1 - General and
Quantitative Analytical Chemistry Lab 1, 2015), Boston University, Boston,
MA.
A thorough understanding of the chemical literature takes practice and finding quality
references that reflect the state-of-the-art in the field can be time-consuming. Having worked
through this chapter, however, you should now be well-equipped to search the chemical literature
with ease, be able to distinguish and evaluate primary references, and properly format a list of
citations skills you will use throughout your scientific careers.
6.4 Annotated Bibliographies
An annotated bibliography is a commented list of sources on a specific topic. It starts with a
short abstract that summarizes the major takeaway points and is followed by the annotated
references. Each reference entry is formatted according to standard reference formatting and is
accompanied by a critical annotation.
As you become more knowledgable and expert in the sciences, the volume of your scientific
library will grow. While the list of papers that you’ve read will start off quite manageable, it
will quickly become crucial that you find an organizational system that will allow you to classify
and comment on each paper. Otherwise, you will find yourself asking the question “now where
did I read that?” not a fun place to be when you go to write a paper.
The idea behind an annotated bibliography is that they become a good way to manage the
references that you’ve read and value. Whether you are citing them in a paper that you are
writing now, or you think that they might be interesting papers to re-read in the future when
you are working on a new project, annotated bibliographies allow you to organize, summarize,
and share the valuable reading that you’ve done.
6.4.1 Content
People who read annotated bibliographies want to gain from the experience of the person who
has read and used the sources. They need a summary of the contents, and they also want a
critical evaluation of the source.
They want to know: the major takeaway point(s) of each source, the source’s intended au-
dience, its relationship to the wider field of research, strengths and weaknesses of the argument,
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 73
and whether it is detailed or broad. In short, the reader of an annotated bibliography wants a
quick and effective insight into some of the sources you have used; either they want to recall
the reason that they found the sources to be useful or they want to get an idea of whether (or
how) they would be useful for their own work. They are, in effect, an executive summary of
the sources.
6.4.2 Language and focus
The language of an annotated bibliography should be formal and objective, following the style of
scientific writing. They are normally written in full sentences, but some use brief and incomplete
sentences when space and brevity are particularly crucial. The focus is on the source (‘the article
presents’, ‘the author concludes ...’), and it is generally written in the present tense.
In an annotated bibliography you are required to outline the type, level, and quality of
the information. Additionally, include the reason for which you find the source to be useful.
Think of yourself as a reviewer of the material. Tell the reader about the text, what it covers,
and in what way. Don’t descend far into the specific details of the work. A sample annotated
bibliography can be found in section 6.6.
6.4.3 Reference managers
It is useful to note that there are also several commercial “Reference Manager” products that
allow you to accomplish this task. Mendeley desktop (free up to 2 GB, all platforms), Papers
for Mac (not free), EndNote (all platforms, not cheap either), and more; the free account for
Mendeley should more than suffice for the work that you do as a student. These managers are
built on the same premise as annotated bibliographies read a paper, categorize it, write the
citation information, leave detailed notes about the paper for later, and make things easy for
storing and distributing this information.
6.5 End-of-chapter Assignment
1. Consider the sample annotated bibliography on the next page.
(a) Based on the theme of the annotated bibliography, use the research tools outlined
in this chapter to find two more articles that would be appropriate to include in this
annotated bibliography. Explain (briefly) how you found the source what tool,
what search parameters, how you refined your search, etc.
(b) For each article you selected in (a), prepare an entry for the annotated bibliography.
Don’t forget to use proper formatting for the references.
(c) Download the full paper for the Grushow entry. Do you agree with the annotated
bibliography entry for this paper? Explain briefly. (You are welcome, and encour-
aged, to read all of the papers they are really interesting.)
2. Use the research approaches outlined in this chapter, and the search tools that you’ve
learned, to find (at least) four papers relevant to your most recent lab.
(a) Briefly describe the approach that you used to find the sources
74 · Undergraduate’s Guide to Writing Chemistry Papers
(b) Briefly (no more than two sentences) explain why you picked these specific sources.
(c) Prepare an annotated bibliography of these sources. Make sure to include an appro-
priate title and abstract.
Writing 6: Research, Scientific Literature, and Annotated Bibliographies · 75
6.6 Sample Annotated Bibliography
Below is an annotated bibliography from my files. It contains some of the more pertinent and
interesting papers that I’ve read on the topic of misconceptions that first-year students have
about how chemical bonds form. Hopefully it will give you some ideas about how annotated
bibliographies can be used effectively to organize your work.
Common misconceptions of first-year students about chemical bonding (2013)
Prepared by B. Abrams, Department of Chemistry, Boston University, Boston MA 02215
For many students, learning chemistry the study of a microscopic world that is beyond their
tangible view proves very difficult. Despite well-meaning intervention from instructors, new
approaches, and updated texts, students leave high school and college chemistry courses with
poor conceptions of chemical phenomena. Chemical bonding is among the most difficult con-
cepts for students to learn properly, without forming strongly routed misconceptions.
In the past, many studies have focused on general misconceptions held by incoming students
in chemistry and the persistence of those misconceptions. Recently, articles and studies have
begun to focus on the misconceptions relating to atomic theory and bonding arguably two of
the more difficult areas for students. A common refrain in these articles is the need for instruc-
tion to be chemically-accurate, and for instructors to be precise and patient in their instruction.
Peterson, R.F.; Treagust, D. “Grade-12 students’ misconceptions of covalent bonding and struc-
ture” J. Chem. Educ. 1989, 66 (6) 459-460
This article presents a diagnostic instrument to assess students’ understanding of covalent
bonding using a two-tier multiple-choice test. The survey of high school indicates that
between 20% and 34% of students demonstrate weakness in topics related to classical
bonding concepts (bond polarity, valence shell, ...). The study concludes that these results
are consistent with the notion that high school teachers may show more concern with
completing topics than helping students to develop authentic comprehension. Overall, this
source is not particularly useful for a discussion of learning of modern bonding concepts,
and is pitched towards a high school or low-level education audience.
Nakhleh, M. “Why some students don’t learn chemistry” J. Chem. Educ. 1992, 69 (3), 191-196
The author argues that many of the problems that students face later in chemistry in-
struction is due to poorly-constructed models that they build early on during instruction.
While this source does not directly discuss bonding, it provides insight into the process
by which the misconceptions about chemistry arise: slowly, over time, and cumulatively.
Two major points made in this paper are (a) about the way students visualize chemical
systems and (b) the need to be precise in definitions so as not to confuse students. While
this paper is not useful for a specific discussion of teaching bonding, it explains a lot about
how incoming students think about chemistry, which is useful for planning how to teach
them.
76 · Undergraduate’s Guide to Writing Chemistry Papers
Grushow, A. “Is it time to retire the Hybrid Atomic Orbital?” J. Chem. Educ. 2011, 88 (7),
860-862
This editorial letter presents a strong argument for eliminating hybrid atomic orbitals
(HAO) from the curriculum. The author suggests two reasons for the persistence of HAO
theory: they help to explain three-dimensional shapes of molecules and predict the orienta-
tions of atoms around central atoms. The author suggests that VSEPR alone is sufficient
to accomplish these goals from a phenomenological point of view, without introducing
fallacies about localized bonding orbitals or degenerate bonds in methane. The major ar-
gument of the paper is that it would be better to relegate discussion of bonding to MO
theory alone, in a course that is appropriate, and not confuse students with a model gives
results contrary to experiment (especially when structure can be reasoned from VSEPR
alone). The paper is well-reasoned and convincing, and is a good reference for support of
major curriculum reform in general chemistry.
Ball, P. “Beyond the bond” Nature 2011, 469 (7328), 26-28
Pitched towards a general scientific audience, this opinion article presents a thorough
discussion of the difficulties associated with the concept of a ‘chemical bond.’ The author
gives a brief historical review of the concept of bonding, including recent complications
introduced by our ability to use ultrashort pulsed lasers to probe chemical motion. The
major focus of this article is to highlight the difficulty (impossibility) of properly defining
a “bond” as a static thing, and that a new notion is needed to replace the long-outdated
view that is commonly held. While the article is informative and authoritative, there is
relatively little substantive discussion of future directions or outlooks in this area other
than to abandon the traditional picture of a covalent bond.
WRITING 7: ADDING MOTIVATION INTRODUCTION
SECTIONS
The Introduction section of a scholarly paper can often be overlooked by the author. Focused
on the innovation that the author is trying to communicate, this very important component can
sometimes be relegated to an after-thought. In reality, while nothing groundbreaking will be
presented in the Introduction, it serves as both the ‘catch’ that grabs the attention of potential
readers and as the foundation upon which the scientific breakthrough is presented. Your paper,
as excellent as it may be, may never actually be read without a good Introduction of course,
your lab instructors will always read your papers in this class.
7.1 Content and Organization
In general, the Introduction is broken down into three specific components:
1. Introducing the research area: the first few lines of the paper are used to generally
describe the research area and to motivate the work. This component is further divided
into three subcomponents:
(a) Identifying the research area: what is this paper going to be talking about (very
general)? This is usually accomplished in the first sentence of the paper. Be careful
not to be too catchy we are writing for an expert audience, not for a general
audience.
Writing Sample 7.1: Identifying the research area
Chlorofluorocarbons (CFCs) are a group of organic compounds, containing
chlorine and fluorine, that have been found to be linked to the depletion of
the ozone layer.
1
This is an example of the first sentence of a paper being written on research
in the general area relating to chlorofluorocarbons. Notice: the actual topic
of the research is NOT presented here. It will be presented in the third
component of the Introduction.
(b) Importance of the research area: why is this area important? What is the motivation
for studying this topic? In general, it will suffice to include a couple of sentences
detailing the reason why the topic is of such high importance. Consider reading the
77
78 · Undergraduate’s Guide to Writing Chemistry Papers
Introductions of other scholarly papers on the same subject to see why other people
consider the area to be significant.
(c) Essential Background about the area: a terse summary of the current state of the
art in the field this should not be exhaustive. Rather, this is used to provide
enough information to give the reader a good understanding of the field. Use relevant
literature sources to explain what has been done in the field to date and explain the
methods that are used in the experiment.
Writing Sample 7.2: Importance of the research area
“In that past, many CFCs were widely used as refrigerants, propellants,
and solvents
1
. The manufacture of such compounds is being phased out by
the Montreal Protocol
2
because it has been shown that CFCs ...”
Continuing the example from above: the importance was already hinted at
in the first sentence (ozone depletion), but is elaborated upon in this sample.
Notice: references must be brought to substantiate any claims made in this
portion of the paper.
2. Identifying a gap: once the field/area has been adequately introduced (discussing the
work already done), we shift towards thinking about what needs to be done. Here, the
author introduces and justifies the need for the body of work in the paper.
It may be difficult to identify a genuine gap (one that actually exists in the current
scientific knowledge) for some of your undergraduate course experiments. This will be
discussed, in detail, later on in this chapter.
Writing Sample 7.3: Identifying the gap
“Although mass spectrometry (MS) achieves the desired detection limits
for ozone
2
, the instrumentation and methodology is costly and time-
consuming.
3
or “The depletion of ozone by CFCs has been the subject of
several studies
4
; however, no studies to date have been reported on the use
of ...”
Both of these samples are examples of a good gap statement. In the first the
gap being described is about a procedure that needs to be improved (made
simpler or cheaper), while the second is an example of an area that needs to
still be explored.
3. Introducing the current work: after the field has been introduced and a gap in the
knowledge has been established, the final component is to describe how this paper will
fill the gap. This is the last paragraph of the Introduction and will typically start with
“In this paper, ...” and will provide a brief description of the general approach that was
Writing 7: Adding Motivation Introduction Sections · 79
used to perform the scientific inquiry. Remember: the use of the first person, or we, is
discouraged in chemistry writing.
7.2 Format and Style
The Introduction should consist almost entirely of prose. Relevant chemical equations can be
included to accomplish one of the goals listed above (motivation, survey of the field, etc.). Sim-
ilarly, figures from the literature or that support the goals can also appear.
Every scientist will eventually develop their own writing style. While there are many areas in
which these can differ, there are some stylistic criteria that are universally accepted. Consider
the following (incomplete) list of considerations for the Introduction section:
You are writing as scientists, not as freshmen students. The objective of the lab is never
to ‘calibrate a pipette’ or ‘learn Beer’s Law’. In other words, the Objective that you wrote
in your lab notebook is unlikely to be the Motivation that you are writing about in your
introduction. See Writing Chapter 1.
By the nature of the types of statements and arguments that need to be made in the
first two components, you will need to provide a decent number of references in the
Introduction. The third component will not require citation unless it is building off of
work that you have already published.
Don’t rely on random facts and certainly never use aphorisms like “it is well-known that
...”. Rather, make sure that your argument (motivation) is focused and supported by
references from the primary (journal articles) or secondary (review articles, book chapters)
sources in the scientific literature. Reminder about internet references: these are to be
avoided, where possible. Recall, however, that journals and books that are accessed online
are not ‘internet’ references and should be cited as journals and books (no URLs).
Don’t be too ‘on-the-nose’ when writing your Introduction. Often, novice science writers
would include statements such as “The objective of this experiment was ...”. Instead, con-
sider a more elegant version, such as “Given the need for, ...” or “Given the importance,
...”. Stylistic elements like these will make your argument stronger, because it will not
sound like you are trying to check-off components from your checklist (though, in reality,
you are).
Remember: less is more. Focus on information density, not on length.
7.2.1 Voice and Tenses
The beginning of the Introduction, especially when discussing the importance of the research
area, is often written in the Present Perfect. The Present Perfect is typically used to indicate
work that was done in the past that is still considered to be true in the present. Notice the
bolded words in Writing Sample 7.2 this is an example of present perfect in the passive voice.
In general, when referring to work done by others (in the past), either the Present or Present
Perfect tenses should be used.
80 · Undergraduate’s Guide to Writing Chemistry Papers
Writing Example 7.1: Proper tenses for citing work
Incorrect:
Abrams et al. demonstrated that CFC levels can be easily ...
Correct:
Recent advances
3
have led to a method for ...
The first version is written in the past tense and is not correct. The second
version is written in the present perfect in the active voice and is a correct
way of citing their work. Additionally, notice the correct use of citations
in the second version. Note: in general, try to avoid mentioning research
groups by name.
The final component of the introduction, describing the work that was done, will contain
both the past and present tenses, where appropriate. The work done in the past should be
described in the past tense (“the structure was determined by ...”), while the truths that are
gleaned from the research are in the present tense (“the structure indicates that ...”).
7.2.2 Length and Depth
It is important to remember to be concise without being terse, and complete without containing
semi-relevant or irrelevant information. Less is more. Do not use more sentences, or overly
verbose clauses, in order to increase the length of your paper. You will not be assessed based
on the length of your paper, rather by its quality and completeness. Moreover, it is completely
inappropriate to write more than is necessary.
Finding this balance can sometimes be hard for a novice science writer. We’ve already run
into this problem with our Results and Discussion sections. You’ll recall that we discovered that
preparing a good, complete checklist before starting to write the paper is extremely helpful
when it comes to being complete and concise.
For the purposes of papers that you will write as a student, you should limit your Intro-
duction to about one double-spaced page
4
(unless otherwise specified) the length should be
commensurate with the amount of information that must be communicated. It is not appro-
priate to have unnecessarily lengthy Introductions. That said, though the page limit may seem
restrictive, it is also important that the Introduction flow well from paragraph to paragraph.
It should not seem like you are simply checking off components from a list; rather, make sure
to utilize appropriate segues in order to connect ideas.
7.3 Argument and Introductions
The Introduction section is the component of the paper in which the author motivates the
science. The goal of the introduction is to captivate the reader on two levels: by demonstrating
the research area is significant and interesting (motivation), and by showing how this specific
study contributes something novel and crucial to the field (gap). Revisit previous chapters to
review Arguments.
4
Though shorter than one page is perfectly fine. Just remember to include the three necessary moves and
sufficient detail.
Writing 7: Adding Motivation Introduction Sections · 81
It should be noted, however, that in undergraduate coursework we will not really have
any true “gaps,” with few exceptions. Since it would be inappropriate to invent a fake gap
statement, we will replace this component with an Objective statement.
Writing Sample 7.4: Gap vs. Objective
Gap: “The depletion of ozone by CFCs has been the subject of several
studies
4
; however, no studies to date have been reported on the use of ...”
Objective: “Given the need to quantify CFCs in atmospheric samples, this
study demonstrates how ...”
The ‘gap’ version indicates that the paper will discuss a novel approach to
the problem, whereas the ‘objective’ version does not.
7.4 Common Mistakes
The most common mistakes that people make when writing their Introductions are:
Writing the Introduction too early! This section should be written last (except for the
Title and Abstract), once it is clear exactly what direction the paper will take. Writing
Introductions too early usually results in unnecessary rewriting.
Using style-inappropriate terms such as: we looked into, we saw, or researchers (these
should never be used). Research may be used sparingly, but is not recommended.
Using the the term current work improperly. In the context of a scientific paper, cur-
rent work refers to the author’s work that is being presented in that paper (not in the
literature).
Presenting results in the Introduction. Remember, this is not an Abstract. The Intro-
duction is for talking about the goals and the current state of the art.
Writing Example 7.2: Introduction (1)
Incorrect:
“Red socks were washed with Downy and Target detergents for
5.5 minutes each at 70
C . Stains were removed 78.5±0.1% of
the time with Downy detergent and 89.6 ± 0.2% using Target
detergent.”
Correct:
“In this work, the removal of chocolate stains on red socks using
a washing machine is investigated. This was done by washing
red socks with a variety of detergents for various lengths of time
using ...”
In the first version, results and unnecessary experimental details are pro-
vided. In the second, the objective was discussed.
82 · Undergraduate’s Guide to Writing Chemistry Papers
Providing too much detail vis-a-vis the experimental details in the Introduction.
Writing Example 7.3: Detailed vs. General
Incorrect:
In this work, the ozone concentration in a 1.0 L gas sample was
determined by aspirating 10.0 µL into an Agilent 1080 GC-MS
with column temperature static at 200
C.
Correct:
In this paper, a method for measuring ozone concentrations in
gas samples via mass spectrometry is presented. The analyses
were performed on a GC-MS equipped with an electron capture
detector (ECD) in order to achieve more accurate results in half
of the time.
There should be little repetition in journal articles. The first version is
appropriate for an Experimental, in which precise experimental details are
provided. For an Introduction, it is better to simply describe the general
approach (second version).
Not citing enough literature sources many Introduction components require citations.
Citations are required any time you present information that is not scientific common-
knowledge.
Using direct quotes. While direct quotations are common in other forms of writing, this
practice is rare and highly discouraged in scientific writing.
7.5 End-of-chapter assignment
1. Review the major ideas in Writing Chapter 5, especially the section on ‘What to avoid in
scientific writing.’ Think back to your old science writing (perhaps from High School or
another science class), specifically to how you would have written an Introduction section,
and suggest two items that you will change or eliminate when you write your Introduction
sections for scholarly papers. Explain very briefly.
2. Prepare a GRA and checklist for your paper (IRDC Introduction, Results and Discus-
sion, and Conclusion). Note: consider following the suggested order below.
3. Use the information in this chapter, the checklist you just constructed, and the details
from previous chapters, when writing your paper (IRDC).
Suggested order for approaching writing a paper:
1. Start by identifying each of the following for your experiment: a scientific motivation for
the experiment, the objective of the experiment, and the educational purpose of the lab.
2. Consider all of the data collected in the experiment. Tabulate the data that are relevant
for this lab and assign them as raw data, processed, or statistic. Indicate where you intend
to include these data in your report.
Writing 7: Adding Motivation Introduction Sections · 83
3. After examining the Results of your experiment, do enough research to make sure that
your satisfied that you understand your results (why you got them? what they mean?).
Prepare a GRA for the Results and Discussion and Conclusion sections. Focus on devel-
oping a strong argument.
4. Based on your GRA and table of data, prepare a checklist for the Results and Discussion,
Conclusion, and Supporting Information sections.
5. Now that you’ve outlined the main part of your paper, it is time to start thinking about
the Introduction. Identify the following components and add them to your checklist:
motivation and gap. Consider using a table (like the one below) for your checklist. Notice
that the table leaves room for the references that you used for each component of the
Introduction.
Required Component: Reference
(1a) The research area is ...
(1b) This area is important because ...
(1c) Background info that needs to be included about area:
(2) The “Gap” is ...
(3a) In this paper ...
(3b) (The general approach was ...)
6. Re-read your checklist to make sure that your paper will make sense and present a strong
argument. Does the Introduction that you’ve planned motivate the argument and claims
that you will make in the Results and Discussion? You are now ready to write your paper.
84 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 8: SHARING YOUR PROCESS EXPERIMENTAL
SECTIONS
The Experimental can be very difficult to write well. There needs to be a balance between a good
amount of detail and readability. The sole purpose of the Experimental section of a scientific
paper is to convey the exact method by which the experiment was carried out. An appropriately-
pitched Experimental will omit the details that any novice would know (such as how to take a
mass by difference, clean a burette, standardize a reagent, etc.) and provide enough description,
observations, and details to sufficiently guide an expert to be able to successfully repeat the
experiment.
Additionally, the Experimental section can often suffer the most from the problems of bad
writing or not being concise. Make sure to review Writing Chapter 5 before starting to write
your Experimental section.
8.1 Content and Organization
In general, the Experimental will not include information that belongs in the Results, such as
spectral data and crystal data. For synthetic papers, yields are often presented in the Experi-
mental.
It is important that the Experimental be complete with the details that another scientist would
require in order to reproduce the experiment. These include:
Exact masses of reagents when they are relevant to the success of the method. In
these cases, report the actual amounts and concentrations that you used; do not use the
approximate amounts like those that would be listed in a lab manual.
Types/purity of reagents. Avoid using brand names; generic names are more appropriate,
unless the specific brand of reagent is necessary for the experiment – this is not likely. In
the event that the purity of the reagents would affect the experiment, make sure to discuss
the purity or origin of the reagent. Similarly, if the reagent needed to be standardized, this
might be included as a brief statement such as “..., which was standardized according to
Harris
1
(where Harris is the author of the recipe for the standardization with reference).
Descriptions of apparatus that are not standard. Only describe an apparatus if it is
not a standard, or well-known, apparatus. Provide the make and model of analytical
instruments such as spectrophotometers, meters (pH, mV), and diffractometers. Do not
85
86 · Undergraduate’s Guide to Writing Chemistry Papers
identify the specific standard equipment (pipettes, flasks, burettes, etc.), since it is clear
by the number of significant figures in the amounts, the context, and convention which
pieces of glassware would be used to dispense solutions (i.e., everyone knows that you use
a pipette to transfer 25 mL of solution).
Key observations. Make sure to include key, or milestone, observations that would allow
someone to know that they are on the right track or have correctly achieved the desired
result. (e.g., “Adding the acid caused the clear, colorless liquid to turn immediately
cloudy.” or “Some small crystals remained at the bottom of the flask, even after heating.”)
Only include hazards and safety concerns that would not be standard protocal or com-
monly known.
Avoid making explicit reference to tasks, steps, and items that are so commonplace that
any well-prepared practitioner would do them.
8.1.1 Special Subsections
In many journals there are two special subsections of the Experimental section: ‘Reagents and
Materials’ (or ‘Materials and Methods’) and ‘Statistical Methods’ (or ‘Data Analysis’). The
Reagents and Materials section is the first subsection of the Experimental and is where all of
the pertinent information about the chemicals (from where? standardized?), instruments, and
techniques can be found.
Writing Sample 8.1: Materials and Methods
Material and Methods. All solvents and reagents were provided by
the Boston University Analytical Chemistry stockroom and were used as
received without further purification. Visible-light spectra were recorded in
quartz cuvettes using Cary 60 UV/Vis instruments with readings at 0.15
nm intervals...”
The materials and methods subsection gives all of the necessary information
for an expert to repeat the experiment.
The Statistical Methods (or Data Analysis) section is the last subsection of the Experimental
and is where all of the post-processing of the data is discussed (Grubbs test, buoyancy scaling,
peak integration, etc.). From here the reader is able to understand how you were able to take
your experimental data and derive the results that you are about to present. Of course, rarely
are sample calculations presented – unless those calculations represent the novelty of the study.
Not all papers need to include these sections. Using these sections appropriately, however,
can make the Experimental section substantially easier to read (and write) and less cluttered.
8.2 Format and Style
The Experimental almost always consists solely of paragraphs. Diagrams of rare, or new,
apparatus should be included as well. In rare circumstances, such as the preparation of a series
of solutions with small variations, a table can be helpful to minimize unnecessary repetitiveness.
Writing 8: Sharing your Process Experimental Sections · 87
For experiments with several parts, the Experimental section of the paper is broken down
into subsections. This practice makes the Experimental much easier to read and it will make it
easier for someone to reproduce the part of experiment in which they are interested.
In order to distinguish these subsections, the first few words of the paragraph beginning
the subsection should be in bold (see Writing Sample 8.1); we do not use “Part ..” in these
headings. There is no limit to the number of subsections that you can use, but they should be
consistent with the number of parts of the experiment you performed (synthesis, analysis, etc.).
Some additional formatting guidelines:
No bullets this is not a procedure in a lab manual.
When using a standard method, cite the literature and do not repeat the instructions.
Variations of literature procedures are cited along with details of the changes.
Include the observations within the procedure when they are relevant or important.
8.2.1 Voice and Tenses
In chemistry, physics, and many biochemistry journals, Experimental sections are written in
the past tense and passive voice. A primary aspect of the argument being made by a journal
article is that the science performed is completely objective and independent of the researcher.
Prose written in the past tense and passive voice gives the impression that the experiment could
be carried out by anyone, and that anyone performing the experiment would have achieved the
same results. You will find it useful to review important stylistic guidelines for numbers, values,
and prose that are presented in Writing Chapter 2.
8.2.2 Length and Depth
There is no maximum length for Experimentals. That said, instructions need to be concise and
to the point. Do not add unnecessary flourishes to make it “sound better”; less is more.
Writing Example 8.1: Sentence Consolidation
Incorrect:
“A 5.00 g sample of Fe(NO
3
)
3
(s) was weighed on a balance.
The solid was transferred to a 100 mL volumetric flask. A 20
mL volumetric pipette was used to add 20.00 mL of 0.1 M nitric
acid to the flask. Water was added to the volumetric flask until
the mark. The final solution concentration was found to be 0.05
M.”
Correct:
“A 100-mL 0.05 M Fe
3+
(aq) solution was prepared from
Fe(NO
3
)
3
(s) diluted with 0.1 M HNO
3
.”
Both statements communicate the exact same procedure. The second ver-
sion, which is much better than the first, is consistent with what we’d
expect to see in a scholarly paper, while the first looks like a Procedure
from a lab manual that has been written in the passive voice.
88 · Undergraduate’s Guide to Writing Chemistry Papers
Equally important is the need to be complete. Do not omit important details in the name of
brevity.
8.2.3 Identifying Equipment
There are four classes of equipment that are discussed in a scholarly paper: standard (make
and model are irrelevant), standard (make and model are important), non-specific, and novel.
Standard (relevant model) equipment refers to instrumentation where the accuracy and
precision will vary from model to model (e.g., spectrophotometers, melting point appa-
ratus). This type of equipment will be referred to by make and model and should be
enumerated in the Materials and Methods subsection.
In the case of Standard (model irrelevant) equipment, while the author does need to specify
the general type of instrument (stainless-steel temperature probe, ice bath), the piece of
equipment does not have to be identified by make and model in the report.
Writing Sample 8.2: Make and model (1)
“The absorbances of all solutions were measured at 510 nm on a CCD Array
UV/Vis spectrophotometer that was blanked with deionized water.”
Every UV/Vis spectrophotometer is built to different accuracy and precision
specifications. Here, it is important to mention the exact make and model.
Notice that this statement concisely describes the spectrophotometric mea-
surements, as well as the relevant parameters: wavelength and blank.
Writing Sample 8.3: Make and model (2)
“After drying for 2 hr at 100
C in a gravity drying oven, the final yield of
product was 0.5250 g (85.55% yield).”
Notice that the equipment used for recording the mass is not mentioned.
There are two reasons: (1) it is proper practice to measure mass using an-
alytical balances when the exact mass is needed; and (2) the four digits of
precision could only have been measured on an analytical balance, which is
something that an expert would know.
Non-specific instrumentation includes balances (all), glassware (all), and thermometers.
It is clear from the degree of precision reported exactly what class of either is used. It is
not necessary to even mention the type of instrument.
Finally, novel equipment is equipment designed specifically for the experiment being pre-
sented. For this type of equipment, detailed instructions and a diagram or figure are
required. Sometimes, when the novel piece of apparatus is not the focus of the experi-
ment, these details are presented in the Supporting Information of the paper.
Writing 8: Sharing your Process Experimental Sections · 89
8.3 Writing a concise Experimental section
The nature of the scientific process (multiple trials, incremental modifications, comparison of
similar data sets, etc.) makes it easy to fall into the trap of repetitive writing. The complicated,
multi-step techniques or analyses can also lead to tendency for verbose descriptions. Finding a
balance between clarity and brevity can be difficult, so you must carefully consider the infor-
mation you are trying to convey, as well as the audience to whom you are writing.
It may help to keep some questions to yourself in mind while writing:
1. “Would the level of scientist who was interested in reading and replicating this work need
me to explain the details of this step, or just the fact that it was done?”: Many techniques
you will employ in the chemistry are routine or commonly practiced, and would therefore
be familiar to the reader. While it is difficult at first to make the distinction (since most
of these techniques are fairly new to you), this ability will develop over time as you are
exposed to more types of experiments and have a chance to see which methods appear
over and over again. (Writing Example 8.2)
Writing Example 8.2: Process Details
Incorrect:
“The powder of sodium sulfate (Na
2
SO
4
) was placed in the 200
mL volumetric flask and then DI water was added to about
halfway, creating a dark blue solution. The mixture was then
swirled around with the water to ensure good mixing. More DI
water was then slowly added until the edges of the meniscus
came up to the mark. The solution was then capped and mixed
by inverting the flask five times. This resulted in a final solution
that was lighter blue. The final concentration was 0.0081 M.”
Correct:
“A 200.00 mL sample of 0.0081 M Na
2
SO
4
was prepared, re-
sulting in a uniform and clear, light blue solution.”
The first version reads like a lab manual written in the past tense. The
second isolates the useful information in a clear and concise manner. Note:
the use of proper significant figures implies the use of volumetric glassware
(this may also have been specified in the Materials and Methods section).
Also, steps like thorough mixing are assumed as they are proper practice.
2. “Was the same exact technique used on all samples?” or even “Was there only that
one systematic difference implemented among the trials?”: If a method was used in the
same way across multiple trials, a single explicit description using one representative set
of numbers is sufficient with the rest summarized as replications. For cases with just
one small accidental difference or a series of systematic differences, these can also be
summarized with a single concise sentence (see Writing Example 8.3). Additionally, as
you develop and learn more expert-like skills, you will come to learn which things need
not be said at all (as they are commonplace and implied).
90 · Undergraduate’s Guide to Writing Chemistry Papers
3. “Does having these data points all written out in long sentences really make my point or
is there a better way to represent it?”: Having multiple sentences with lots of numbers,
or many sentences outlining multiple differences (e.g., solution preparation) is confusing
and can contribute to your point being lost. These cases are better served by a table (or
occasionally a figure).
Writing Example 8.3: Repetition
Incorrect:
Preparation of solution A. Unknown tablet A (1.0561 g) was
crushed and dissolved in 50.00 mL of DI water with continuous
swirling over heat. A 5.00 mL aliquot of this solution was then
transferred to a beaker and 5.00 mL of 0.1 M HNO
3
(aq) was
added. . . .
Preparation of solution B. Then unknown tablet B (1.1201 g)
was crushed and dissolved in 50.00 mL of DI water with contin-
uous swirling over heat. A 5.00 mL aliquot of this solution was
also transferred to a beaker and 5.00 mL of 0.1 M HNO
3
(aq)
was added. . . .
Correct:
Preparation of samples. Tablets were dissolved in 50.00 mL of
DI water with continuous swirling over heat. A 5.00 mL aliquot
of this solution was then transferred to a beaker and 5.00 mL
of 0.1 M HNO
3
were added.. . .
In the first example, the nearly identical procedure was described twice
(though the mere act of separating them is already an improvement over
an Experimental written in one, long section). In the second example, the
two preparations are combined.
8.4 Common Mistakes
The most common mistakes that people make when writing their Experimentals are:
There must be a space between values and units (and zeros before decimal points).
Writing Example 8.4: Units
Incorrect:
“A .2021 gram sample of NaCl(s) was added to 15mL of ...”
Correct:
“A 0.2021 g sample of NaCl(s) was added to 15 mL ...”
Never spell out units, and there should be a space between 15 and mL.
Writing 8: Sharing your Process Experimental Sections · 91
Remember that it is important to not be too vague. Important details must be included.
Writing Example 8.5: General vs. Specific
Incorrect:
“A bromine solution was added to the flask”
Correct:
“When 10.00 mL of bromine were added to the reaction flask,
the solution immediately turned bright orange.”
The first statement lacks quantization and specificity. How much bromine
was added? To what flask? What happened when you added it?
A big mistake is that students often write their Experimental as if it were a lab manual:
Writing Example 8.6: Experimental vs. Procedure
Incorrect:
“Add approximately 0.2 g of NaCl(s) to a volumetric flask.”
Correct:
“A 0.2021 g sample of NaCl(s), diluted to 100 mL, ...”
The first statement is more reminiscent of a lab handout instruction; it is
written with approximate amounts and as an instruction (do this). The
second version correctly reports the precise amount of solid used and does
so in a passive voice.
One of the most substantial problems is when someone reports too much raw information,
such as masses of weighing jars and paper:
Writing Example 8.7: Raw Data
Incorrect:
“A sample of NaCl(s) was weighed out in a weighing jar with
mass 15.1000 g. The mass of the jar and NaCl was 15.3041
g. After delivering the NaCl to the reaction flask, the jar was
reweighed at 15.1020 g. The mass of NaCl delivered was 0.2021
g.”
Correct:
“A 0.2021 g sample of NaCl(s) was added ...”
The first version is verbose, clumsy, and unnecessary. The second version
correctly, concisely, and properly reports the precise amount of solid used
(an expert knows to record mass by difference).
92 · Undergraduate’s Guide to Writing Chemistry Papers
8.5 Past experiments in context
It can be very difficult, especially for novices, to get a good feeling for the level of detail that is
expected in the Experimental section of a scholarly paper. Often, they include way too much
detail about the work that they’ve done. As these courses are usually their first foray into
real-world experimental science, this is not unsurprising.
To help illustrate, let’s consider an example from a lab that is common in first-year chem-
istry courses: titrating weak acids. Labs of this nature have two main parts: titrating NaOH
solutions with a primary standard (like KHP) in order to determine the precise concentration
and using the standardized NaOH to determine the amount of an unknown weak acid (KHP,
for example) in an unknown mixture. The entire Experimental could be written as: “Samples
of KHP-containing unknowns were titrated against standardized 0.1 M NaOH to a phenolph-
thalein endpoint.” The first part of the experiment is replaced with the word ‘standardized’
any expert should know how to standardize NaOH, or they could purchase pre-standardized
base. Additionally, the exact masses of the samples are only nominally relevant since we are
determining the concentration, which is an intensive property (one that is sample-size indepen-
dent). Finally, number of measurements are not included in the Experimental.
A good approach for learning the appropriate balance in an Experimental is to look at similar
types of papers in published journals, and mimicking the level of detail and the language used.
8.6 Preparing to write an Experimental Section
The Experimental section of a paper is the only section that truly ‘stands alone.’ While the
strength of the methods used often contribute to the strength of the argument, it is rarely
the case that the methods are a cornerstone of the Argument being made. The most glaring
exception to this general idea is when the paper is focused on the development of a new method.
As a result, the Experimental can usually be written in parallel to the rest of the paper.
It is challenging to sort out what details are important, especially if you are trying to do so
as you write. A suggestion to help guide you in your preparation of your Experimental sections,
is to use a table to help sort out the different subsections that you will use, and help determine
what details are necessary to include and which details should be omitted.
To illustrate the power of this approach, let’s consider an experiment to determine iron
content by AAS. A suggested breakdown of part of the Experimental is tabulated below.
Subsection: Necessary Information Unnecessary Information
Materials and - Glassware prep for trace - Specific glassware used
Methods metal analysis (soaking, - Source of chemicals
and chelating solution) (other than the cereal studied)
- Make/model of AA
Cereal digestion - Brand and mass of cereal - Glassware used
- Volume of acid (and which acid) - Exact concentration of acid
- Heat for 30 minutes a low boil - Number of washes
- Filter to remove all solids - Number of filtrations
Preparing standard - Volumes of digestion - Glassware used
addition solutions - Range of concentrations of stock - Exact concentration of Fe stock
Writing 8: Sharing your Process Experimental Sections · 93
8.7 End-of-chapter assignment
1. Consider the following excerpt from a not-so-well-done Experimental section:
To titrate a solution of sodium oxalate: A burette was filled with .02 Molar KmnO4.
After it was found that no air bubbles were trapped in the tip of the pipette, the initial
volume was recorded. Using a weighing bottle, 0.2315g were weighed out. Fifty mL of 4.5
Molar sulfuric acid, 200 mL of distilled water, and the sodium oxalate were added to a
400 mL flask. The solution was stirred until all of the solid dissolved. Approximately 30
mL of KmnO4 was added to the solution. The solution turned dark purple. The solution
was then continued to be stirred due to the fact that the titration was incomplete. Once
the solution becomes clear, place it on a hot plate in a fume hood and heat it until it’s
temperature was between 55 and 60
C. While the solution was still warm, the titration
was completed by adding KmnO4 until a significant amount of pale pink color persisted,
and the end volume was then recorded. The data were recorded in the lab notebook.
(a) Using the information in this chapter, identify at least five issues with the style, for-
mat, language, or chemistry of the Experimental section excerpt above. For example:
.02 should be written as 0.02.
(b) How much of this information is really necessary to be included in an Experimental?
Explain your choices briefly.
2. Reproduce the table below on your computer, and use it as a template for a checklist
for an Experimental for a recent lab. The two special subsections and one additional
subsection are shown use as many subsections as you need or deem necessary.
Subsection: Necessary Information Unnecessary Information
Materials and
Methods
Data
Analysis
3. Use the checklist you just constructed when writing your Experimental section for the lab
you chose.
94 · Undergraduate’s Guide to Writing Chemistry Papers
WRITING 9: GETTING YOUR WORK FUNDED RESEARCH
PROPOSALS
The previous chapters have all focused on the development of a single form of scientific writing:
the journal article. There exists, however, a plethora of different forms of communication that
a scientist must master. In this chapter we will learn how to write a Research Proposal. The
main goal of a Research Proposal is to convince someone to give you money. To that end, you
will have to be an expert salesperson and convince your audience, usually a panel of experts in
the field or a closely-related field, that your body of research is worthwhile.
Traditional Research Proposals can be anywhere from 5–50 pages long, depending on the
funding agency and the amount being requested. The Proposal is broken down into four parts:
1. Cover letter: typically a single page in which you introduce yourself to the funding agency
and introduce the topic.
2. Project Summary: a one or two page summary of the project.
3. Project Description: a complete breakdown of the project that includes the background
information, complete enumeration of instrumentation and techniques, timeline for the
project, budget, and expected results. This is the component that determines the length
of the proposal.
4. References: a complete set of properly-formatted references for the proposal.
The proposal that you will be writing will be the Project Summary with References (#2 and
#4). As a side note: the Project Summary that you are working on here is exactly the type
of proposal that students applying for summer fellowships and research internships might be
asked to write.
9.1 Project Summary
In many ways, the Project Summary is the most important part of any research proposal. The
Project Summary has the longest shelf-life of any submitted report or proposal; in fact, the
funding agencies will retain the Project Summary, as record of the project, even long after the
project has been completed. It is for that reason that the Project Summary must be written
as if it were a stand-alone document.
95
96 · Undergraduate’s Guide to Writing Chemistry Papers
9.1.1 Content and Organization
The Project Summary has five major components (the first three components are, traditionally,
heavily cited):
1. Introduce Topic: The proposal should start by introducing the topic of the research project
and discussing the general project goals. This should be fairly general, as the specific
expected outcomes will be discussed later in the proposal.
2. Emphasize Significance: After introducing the topic and goals, the overarching significance
of the research is discussed. Why is this important? What are you hoping to demonstrate?
What effect will this have on the field and on the world? It is very important to be
convincing here. Remember: your goal is to elicit carefully-guarded, highly-coveted, and
difficult to acquire funds.
3. Summarize Methods: This is the most important (and hence longest) component of the
proposal. Now that the topic has been thoroughly introduced and motivated, the next
step is to summarize the methods that will be used in the context of the goals (aims) of
the work. Here, Methods refers to (a) traditional techniques that will be used, (b) non-
standard equipment/instrumentation that is needed, and (c) novel approaches/techniques
that will be developed for the research project. Also discuss and mention if your work
will be some sort of modification or combination of existing techniques.
It is important to be very clear and specific about how the research will be carried
out; this is an integral part of convincing the funding agency that it is worthwhile to
invest in your project. That said, you should not turn this into an Experimental or
procedure. Include details such as specific approaches (digestions, sample preparation,
sample handling, etc.), number of trials, and instrumentation; but refrain from giving
exact amounts, lists of glassware, and step-by-step experimental details.
It is helpful to break up each of the goals (aims) of the project and to list the methods
involved for each. These are often referred to as Specific Aims and are usually part of
a standalone, one-page document. The methods attributed to each of the aims should be
described.
Writing Sample 9.1: Specific Aims
Specific aim #1 is to determine the efficiency of green washing machines
(GWMs) using a combination of digestion techniques and high-performance
liquid chromatography. A series of ten dirty, chocolate-stained sock samples
will be digested according to the Villar method
1,2
.”
Notice that the specific goal is identified first, along with the general ap-
proaches that will be used. Then, the specifics for these methods are given;
enough to give the relevant details of how the work will be carried out, but
not so much as to constitute an experimental procedure.
4. Expected Outcomes: While the actual results of the research are never known a priori,
it is important to have a good sense of what the results might be. Often, authors make
Writing 9: Getting your Work Funded Research Proposals · 97
reference to preliminary data or previous work that is similar to the project. If you already
have some preliminary data from your project, this would be a good place to use that
work.
5. Summarize Broader Impacts: The final portion of the Project Summary is to summarize
the broader impact that the results of the research will have on the field and, possibly, the
world. It is important that these outcomes relate directly to the motivations, or at least
be framed in the context of the motivations already mentioned earlier in the proposal. Do
not go overboard in suggesting fantastical impacts or those that will remain far into the
future, even after the project is complete.
9.1.2 Format and Style
The project summary always starts with the tile of the proposal, the name(s) of the principal
investigators (PIs), the institution(s) of the PIs, and the date of the proposal. Some funding
agencies assign codes or identifiers to PIs and should be included when relevant.
In general, Project Summaries are no longer than two pages with size-10 font. Spacing
and margins are rarely specified in the requirements. In general, project summaries and short
proposals are not broken down into subsections.
Many have the good practice of using bold face for the key words that start one of the above
components. For instance, The goal of this study”. This practice calls the reader’s attention
to these components.
9.1.3 Voice and Tenses
Unlike the journal articles that you’ve written until now, which are mostly written in the past
tense, Research Proposals are partially in the present tense and part in the future tense. All of
the introductory material (components 1 and 2) are written in the present tense, as they discuss
the current state of the scientific art. The methods, expected results, and broader impact, are
all referring to research that may, or may not, take place in the future.
Another major difference between Journal Article writing and Proposal writing is with
regard to the voice. The vast majority of a journal article is written in the passive voice in
order to emphasize that the results obtained are not dependent on the scientist, or research
group, involved anyone would have received the same results. In contrast, in a Research
Proposal only the methods and impacts sections are written in the passive voice (depending
on the context, methods can sometime be in the active voice as well). The remainder of the
proposal is written in the active voice. Unlike in a journal article, where the emphasis is on
the results of the experiment with the idea being that any competent researcher would achieve
the same results, in a Proposal you want to emphasize your own involvement in the successful
outcome of the research you don’t want a funding agency believing that just anyone could do
this research. You want them to believe that it has to be you.
9.1.4 Argument
In the past, your arguments have focused on the results of your experiment and their impli-
cations. Your goal was to convince the reader that your results have substantial impact and
importance in light of the significance of the research area.
98 · Undergraduate’s Guide to Writing Chemistry Papers
The argument you are trying to make in your proposal is quite different. Instead of putting
the results into context, your job is to convince the referees (people that are reviewing your
proposal and deciding on funding) that you will make excellent use of their research funds.
To do this you will need to (a) demonstrate the importance of the research area; (b) explain
your well-thought-out plan that you will use when executing the research; and (c) clarify the
research outcomes and impact that your work will have. Notice the declarative voice this is
deliberate; your proposal should be written with confidence and purpose, with the references
and research to back it up.
Consequently, the Argument that you will write for a proposal will be different than one for
a journal article. Use the above-mentioned guidelines while contemplating your arguments.
9.2 References
This section always starts a new page. The References are listed by number, in order of first-
citation in the Proposal, and are always formatted according to the guidelines and standard
format of the group to which you are submitting the proposal. When page limits are imposed
for research proposals, the References are almost never included in those limits (this will be the
case for us too).
9.3 Common Mistakes
The most common mistakes that people make when writing a Research Proposal are:
Starting to write without developing an Argument and a checklist is a big mistake. The
success of the proposal will be determined by the strength of the argument, the merits
of the science, and the quality of the writing. It is critical that a good checklist, which
is designed to form a good argument, be constructed prior to writing the proposal. A
successful checklist will also indicate which the corresponding component of the Project
Summary for each item on the checklist and will also include references.
Often, proposals are written in collaboration with other researchers. It is a mistake to
simply split up the work for the proposal without first agreeing on a shared vision for the
proposal. This inevitably leads to a disjointed proposal and duplication of effort.
While most reviewers might not focus on them explicitly, using the wrong verb tenses (or
moods) makes the proposal a more amateurish read. The general feeling that the reviewer
gets from reading the proposal can often make the difference between getting funded, or
not.
Make sure that the broader impacts are directly related to the significance of the research,
as this will help to tie together the whole proposal. Also, be reasonable about your
expected impacts. Your undergraduate lab will not even come close to curing disease, so
don’t imply that in your impacts.
While it is difficult for first-time proposal writers, it is important to reach an appropriate
balance of detail in the methods section.
WRITING 10: MAKING AND DELIVERING EFFECTIVE
RESEARCH PRESENTATIONS
10.1 Ten guidelines for constructing an effective presentation
Presentations, like journal articles, are a vehicle for communicating the results of scientific
inquiry – an argument. Before planning to prepare a research presentation, review the material
and suggestions in previous chapters on how to approach crafting your argument. Specifically,
you will need to make sure that you’ve completely analyzed all of your data, outlined your
argument with a GRA, and prepared a strong argument. Like in written papers, the effectiveness
of a presentation is limited by the quality and strength of the argument that you intend to make.
Consequently, make sure to fully flesh-out the argument of your work before starting to prepare
your presentation (or paper) – this is the only way to ensure that you will end up with a strong,
persuasive presentation.
Once you’ve settled on the argument that you intend to make, you can start to think
about organizing your work into a paper or a presentation. The content being presented, the
organization of the materials, the level of detail presented, the aesthetic quality of the slides, and
the prose of the speaker are all important aspects that go into crafting a successful presentation.
Consider the guidelines in this chapter when preparing your presentations.
10.1.1 Think message, not slides
When you start by thinking “what’s on slide 1, what’s on slide 2,” you’re in trouble, unless you
want to end up with a lot of boring slides that say nothing. Before you start on any slides,
figure out the answers to the following questions. Ask yourself: what point do I really want to
make to my audience? What message do I want to plant in my audience’s brains? Now imagine
that message on a billboard You will make a slide to convey that specific message. Together,
these slides will unfold the strong argument that you want to make. Revisit all of the details
presented in Writing Chapter 3 about making a strong argument.
Consider the examples of good and poor points that are presented in Table 10.1. Which
kind of presentation would you rather go to? The one that makes some kind of point, or one
with no point, no focus and sometimes, no end?
99
100 · Undergraduate’s Guide to Writing Chemistry Papers
Table 10.1: Examples of good and bad presentation points.
Good points Poor points
Why method X is so good at measuring ... Here is method X
Using chemical A increases efficiency by 78% We used chemical A
The advantage of this method is ... Information about a method
10.1.2 Decide why your audience should care
Brainstorm all the possible reasons your audience should care about your point. Imagine that
you are at a conference of professional and academic chemists. How can you present your
material so that they do not feel that the 15 minutes they spent listening to your talk was not
wasted time? Find a reason why the audience members should be interested in your particular
topic. This is a lot like the “motivation” of a journal article or communication. Start by
brainstorming asking yourself what interests you about your topic. Brainstorm means “come
up with a lot of possible answers.” You should have at least three answers written down on
a list to ensure you’ve really thought about your audience. Remember that the broader your
reasons, the more people will be interested in your presentation. Consider some examples:
You are introducing a topic that is not currently discussed or well-known
You are going to add understanding of X to what is already known
You have tested and improved on a new methodology that is useful
You have learned a new property of a material that is interesting
10.1.3 Make points, don’t cover topics
Once you have the main message that you want to convey, you can divide your presentation
into sections. But instead of thinking of sections as a generic category such as introduction,
history, conclusion think of all the points you want to make that will contribute to your main
message (the one on the billboard). Identifying these points makes a big difference how effective
your presentation is at engaging your audience. Note: you’re still not writing/making slides.
Once you have these sub-points, those slides will practically design themselves. See table 10.2
for examples of some good sub-points.
Table 10.2: Examples of good and bad sub-points.
Good sub-points Poor sub-points
Explain the problem and its relevance (what State how you personally became
we know and why it matters) interested in or chose this topic.
Relevant methods and results with analysis Description of every mL measured
Your important findings and brief analysis of Your findings with no analysis
their relevance
Deeper analysis of the important findings and Conclusion with no take-home message
future development
Writing 10: Making and Delivering Effective Research Presentations · 101
10.1.4 Tell a story have a flow
Once you’ve identified good sub-points, it’s time to determine the best sequence for them. Try
to imagine what would make the most convincing argument and what would be the most inter-
esting flow for your audience. Doing this well means you avoid having presentations that are
just ‘one thing after another.’ Instead, they will have a compelling thread that ties the whole
thing together a cohesive argument.
Consider the following sequence for your presentation (argument):
1. start by stating the problem (motivation),
2. then discuss the most important methods that were used to study the problem,
3. then present the solution to the problem that you have uncovered in your research (re-
sults),
4. then discuss the impact and consequences of the results,
5. remind the audience of the important findings and their broader impacts (refer back to
#1),
6. suggest possible future research avenues or developments that may come from your re-
search, and
7. thank the appropriate people (advisors, funding agencies, and collaborators) at the end.
Notice that this sequence very closely matches the same argument sequence that we’ve developed
for our written work.
10.1.5 Get creative
Think of figures, metaphors, examples, and analogies as a way to back up your sub-points and
make them ‘sticky’. Here, sticky means that the sub-point sticks with the audience you because
it came wrapped up in a story, an image, or an analogy. Sure you have data, and you need
that data, but persuasion doesn’t depend on logic and data alone (if it did no one would smoke
cigarettes or pay $2.50 for a bottle of ‘vitamin water’). What images, symbols, or other visuals
come to mind that might make a stronger impression than just words?
10.1.6 Prepare a presentation “storyboard”
Okay, now you can think in terms of slides! Create a new PowerPoint file using a layout that
includes a title placeholder. Figure out what you want your slides to do and just jot that in the
title placeholder that way, you can see that when you look in the outline tab on the normal
view of presentation or when you look at slide sorter. You’re not designing slides yet, you’re
making a flow and checking that each slide has a purpose a message that the slide is going to
communicate.
102 · Undergraduate’s Guide to Writing Chemistry Papers
10.1.7 Put a real headline on each slide
No one word headings ever. Your heading can be a question or a statement, but it must
say something meaningful. For inspiration, look at the newspaper news, sports, editorials,
no matter where you look in a newspaper, headlines make statements. Having slide titles that
make a point will make your presentation better than just about everyone else’s.
Let’s imagine you’re working on a presentation that explains why copper is unique in giving
high-temperature superconductivity in layered perovskite oxides. Assuming that you have six
slides following the sequence detailed above, your headlines might look like the ones in Table
10.3.
Table 10.3: Examples of good and bad slide headings.
Good headlines Poor headline
Uniqueness of copper in high-T superconductivity Introduction
Z method is most accurate for studying Methods
Cu superconductivity
Copper shown to exhibit X property under Y condition Results
X property increases inversely to heat Graph of X vs. heat
X property disappears in the presence of Fe Figure 2
Y conditions should become standard when creating Conclusion
copper superconductors
This research has been made possible by the following: Acknowledgements
10.1.8 Use the body of the slide to best deliver its message
Ignore your first impulse to have a nice bulleted list of one word items. That might be your
second and third impulse too, so resist. Sometimes you do need to use a bulleted list, just
like sometimes you need to use Febreeze instead of actually washing your clothes, but it’s an
emergency measure.
Slides aren’t supposed to be your teleprompter, though for some presenters that is what it
seems they have become. Slides are supposed to be a visual reinforcement of the speaker’s point.
They don’t have to say everything. In fact, they should not. They just have to communicate
one message. Under no circumstances should everything that you intend to say be on the slide.
The slides are your prompts, not your script.
Maybe you’ve got an image from your creative period back in #5 use that. Or maybe
you just want to have a few words on the screen but never write every word you’re saying.
Whatever will convey that message best in the quickest, easiest, most impactful way do
that.
10.1.9 Use tables and figures efficiently
Do not include every graph or table that you produced. Figure out what information is necessary
to convince your audience of the point you are making. Make the figures large and legible, and
only include the relevant information. Your viewer is not going to have time to analyze your
tables and graphs in detail, so only present the material that they can absorb in a 1-2 minute
slide. Quickly flipping through several slides of tables or graphs does not help you at all. One
or two good figures to provide the evidence you need to support your claim is all that you need.
Writing 10: Making and Delivering Effective Research Presentations · 103
Do not use “AutoShapes” or excess animation; these elements are distracting and take focus
away from the presenter and, more importantly, from the science. That said, do not make your
slides boring and monochromatic. There is a balance between cheesy and boring.
Add annotations to figures and graphs to highlight important features that you will discuss
and draw the audience’s attention to them. On slides with spectra data, include structures and
equations, where relevant.
10.1.10 Work through steps 1-9 again; Check the balance
It is important that you make several iterative edits to your presentation. Review rules 1-
9 to make sure that you have developed a cogent, organized, and persuasive presentation.
Additionally, make sure to review the list of common mistakes (below) as you work through
your slides. Make sure that each item (slide, table, figure, list) has something meaningful to
add. Tell a story; don’t simply cover topics.
Finally, it is crucial that there be a proper balance to your presentation. Make sure to spend
the vast majority of the time on the actual scientific study the was performed: methods, results,
and implications. After all, that’s why they’ve come to hear your talk. It is not necessary,
nor appropriate, to spend any substantial amount of time on the motivation (they already
listening to the talk and the results should explain why it’s important) and future directions
(you haven’t done them yet!). For a short (15-20 minute) talk, methods and results/discussion
should occupy all but a couple of minutes; longer (45-60 minute) talks will usually spend a little
longer discussing motivation.
10.2 Common mistakes
Be careful of too much white space (not enough information), but don’t make your slides
so cluttered that someone gets lost. Too many words on a slide will require the audience
to spend too much time reading and less time listening to you. In the same vein, a bulleted
list must have at least three points. A table must have more than two columns or rows.
Make sure to use appropriately sized fonts: the defaults in Powerpoint are way too large.
The words should be large enough to be seen, but not so large as to make the slide four
lines.
References should be included on the slides on which they are relevant. They are placed
at the bottom of the slide and in a significantly smaller font than the regular text (still
large enough to read). References must be explicitly mentioned by the presenter. There
is no References or Bibliography slide in chemistry presentations.
Use color if possible, and especially if you are trying to emphasize something. No yellow
or green, though: they never show up properly on a large screen. Only use white slide
backgrounds.
Figures and tables should be complete, clear, easy to read, and of high resolution. Make
sure to follow the guidelines set forth in Writing Chapter 2. You do not need to include
captions, but legends and labels need to be legible. Also, do not copy/paste from PDF
files, this leads to low-resolution images; rather, include these figures as JPEG, PNG, etc.,
and then resize.
104 · Undergraduate’s Guide to Writing Chemistry Papers
For a 15 minute talk you should have approximately 8-12 slides (excluding the title slide
and the slide at the end with the acknowledgements). Your slides are not your presenta-
tion! They accent and provide visual cues for your presentation. Check the balance of
your talk don’t dwell on trivial aspects.
No cue cards. Your slides should give you all of the ‘cues’ that you need. That said, do
not memorize and regurgitate a speech. A good presentation will respond to the slides
and interact with the audience.
Dress respectively for presentations. You do not need to wear a suit, but it is impor-
tant that you be dressed in a way that is respectful and appropriate for an academic or
professional setting.
10.3 End-of-chapter assignment
1. Outlining your presentation.
Prepare a list of points that you intend to make in your presentation (see section
W10.1.1).
Prepare a list of sub-points that will become your presentation (see section section
W10.1.3). Label them as ‘A’, ‘B’, ‘C’, ..., for reference in subsequent questions.
Identify the creative elements (see section W10.1.5) that you intend to use. Assign
each element to a sub-point.
Organize the sub-points in a logical order of slides, and pick a headline for each slide.
Assemble these slides into an outline for your presentation.
2. Slide-deck for the presentation. Use the answers to the above and the guidelines provided
in this chapter to prepare a well-organized presentation for your research project. Include
a PDF copy of the slides in your submission.